How To Cite This Thesis: University of Johannesburg
How To Cite This Thesis: University of Johannesburg
o Attribution — You must give appropriate credit, provide a link to the license, and indicate if
changes were made. You may do so in any reasonable manner, but not in any way that
suggests the licensor endorses you or your use.
o NonCommercial — You may not use the material for commercial purposes.
o ShareAlike — If you remix, transform, or build upon the material, you must distribute your
contributions under the same license as the original.
Surname, Initial(s). (2012) Title of the thesis or dissertation. PhD. (Chemistry)/ M.Sc. (Physics)/
M.A. (Philosophy)/M.Com. (Finance) etc. [Unpublished]: University of Johannesburg. Retrieved
from: https://ujcontent.uj.ac.za/vital/access/manager/Index?site_name=Research%20Output (Accessed:
Date).
Design and Evaluation of a Single Stage Axial Flow
By
Radeshen Moodley
200501035
MAGISTER INGENERIAE
In
MECHANICAL ENGINEERING
At the
At the
UNIVERSITY OF JOHANNESBURG
And
Declaration
I Radeshen Moodley hereby declare that this Master’s Thesis is wholly my own work and has not been
submitted anywhere else for academic credit by either myself or another person.
I understand what plagiarism implies and declare that this Master’s Thesis is a collection of my own ideas,
words, phrases, arguments, figures, results and organization except where reference is explicitly made to
another’s work.
I understand further that any unethical academic behaviour, which includes plagiarism, is seen in a serious
Abstract
This thesis involves the preliminary design and aerodynamic evaluation of a single stage axial flow turbine
rotor for a micro turbojet engine. The design process involves a detailed cycle analysis using Gasturb 12
cycle evaluation code as well as engine testing for a comparable micro turbojet engine. One of the
cornerstone objectives was to obtain reasonable correlation between the test engine data and the Gasturb
12 cycle code. The cycle was evaluated at various operating conditions for the test engine. These would be
experimental test conditions during the engine testing, design point at sea level static condition and
maximum RPM at sea level. The Gasturb code was used to emulate engine test results by showing that the
experimental conditions could, with good accuracy, be recreated in the theoretical software environment
governed largely by empirical methods employed on the analysis of large gas turbines.
Aerodynamic design of the turbine blades was conducted in the Concepts NREC Axial and AxCent design
software environment and compared to a meanline hand calculation for blade geometry. Rotor disk stress
was evaluated at design point operating conditions with aerodynamic loading and free rotation. The engine
test rig at the University of Johannesburg was re-commissioned, calibrated and operated to determine
practical cycle operation and design point behaviour of a micro turbojet engine. Characteristics of the engine
at the three evaluation conditions were determined. These conditions were run in Gasturb 12 and the results
compared to experimental data from a similar micro turbojet engine used for experimental data collection.
Evaluation of mechanical aspects of the turbine design will follow in future work along with complete rotor
and NGV integration into the turbojet used to collect the baseline data. Due to resource limitations a
computational fluid dynamics study (CFD) of the rotor and NGV design was only completed. The
manufacture and testing of the rotor and NGV did not take place. The designed turbine rotor was not
manufactured or integrated into an engine for testing due resource restraint and time restraints.
Acknowledgements
The student would like to thank Prof. Japie van Wyk for his invaluable support, dedication and continuous
encouragement throughout the many years that resulted in this document. For your many hours dedicated
above and beyond the call of duty, thank you! Additional and gracious thanks to:
Mr D. Madyira, for bringing this transcript to life with the hard hitting questions and suggestions,
thank you!
Mr J.G. Benade, for the help and advice both technically and aesthetically in all aspects of the thesis
and especially for asking all the hard questions when they were needed most!
Mr P. van Deventer, technical and practical advice on the test rig, engine specification and
especially regarding engine start procedure, engine specification, and engine controller setup.
Ms N. Janse van Rensburg, for her assistance during presentations at the University MERF
Prof. I. Sigalis, University of the Witwatersrand, for his guidance and forthcoming approach to the
Prof. R. F. Laubscher, for his assistance in developing my understanding of titanium aluminides and
structural ceramics.
Prof. A. Nurick, for his lengthy discussions regarding the heat transfer implications of the work.
Dr P. F. J. Henning (UJ HOD Mechanical Engineering 2011-2014), for advice, guidance, pressure to
Dr G. C. Snedden, my boss at the CSIR and my friend at the office, for his endless advice,
encouragement and understanding of turbine design, trials and tribulations, and support in
Dr D. Dunn my colleague and friend at CSIR, for breaking down the ins and outs of CFD and its
limitations, for the endless chats about life the universe, engine braking and everything else.
My family, Mom, Dad, Brother and Girlfriend (Ronelle), for all the support and understanding during
Table of Contents
Declaration 2
Abstract 3
Acknowledgements 4
Table of Contents 5
List of Figures 7
List of Equations 9
List of Tables 11
Nomenclature 12
Abbreviations 14
1 Introduction 15
1.1 Background 15
1.2 Aim 16
1.3 Objectives 16
1.4 Scope 16
1.5 Methodology 17
2 Literature Review 19
2.1 Background Remarks 19
2.2 A Brief History of the Gas Turbine in Aircraft Propulsion 19
2.3 The Simple Turbojet Cycle 20
2.4 Turbine Operation and Design Theory 25
2.4.1 The Role of the Turbine 25
2.4.2 Radial and Axial Turbines 25
2.4.3 Impulse and Reaction Turbines 28
2.5 Micro Turbojets 30
2.5.1 Applications and Layout 30
2.5.2 Design and performance limitations 31
2.5.3 Total Pressure Loss in an Axial Turbine Stage 32
2.6 Gas Turbine Metallurgy 34
2.6.1 Nickel Super Alloys 34
2.6.2 Manufacturing Techniques 35
2.6.3 Structural Ceramics in Turbine Design 37
2.7 Concluding Remarks 38
3 Experimental Test Programme 39
3.1 Background Remarks 39
3.2 Limitations 39
3.3 Aim of the Experimental Work 40
3.4 Experimental Procedure 40
3.5 Test Rig and Data Acquisition System 40
3.5.1 Pressure Measurement 40
3.5.2 Temperature Measurement 41
3.5.3 General Measurements 41
3.6 Sensor Calibration 41
3.7 Experimental Results 42
3.8 Concluding Remarks 45
4 Cycle Analysis 47
4.1 Background Remarks 47
4.2 Aim 47
4.3 Component Efficiencies and Cycle Analysis 48
4.4 Methodology 51
4.5 Cycle Sensitivity Analysis 51
4.5.1 Compressor Pressure Ratio and Isentropic Efficiency 52
4.5.2 Turbine Inlet Temperature and Isentropic Efficiency 53
4.6 Cycle Analysis using Hand Calculation Methodology 56
List of Figures
List of Equations
List of Tables
Table 4-1: Gasturb cycle results summary 68
Table 5-1: Station calculation results across the turbine stage 77
Table 5-2: Summary of hand calculations at mean radius and design condition 77
Table 5-3: Prismatic blade turbine performance map results 82
Table 5-4: Twisted blade turbine performance map results 85
Table 5-5: NREC stage thermodynamic summary 88
Table 5-6: Concepts NREC optimized blade geometry for a throat Mach number 0.75 88
Table 5-7: Calculated mean line blade angles for throat Mach number 0.7 89
Table 5-8: Boundary conditions for CFD initialization 90
Table 5-9: Design Objective and Demonstrated CFD Results 95
Table 5-10: Summary of turbine design data, CFD data, rig test data for the turbine and Gasturb
approximation of rig test data at altitude 104
Nomenclature
A Area
Cp Pressure Coefficient
Cx Axial Velocity
Hz Hertz Frequency
K Kelvin Degree
N Newton Force
P0 Total Pressure
P Pressure
r Radius
T Temperature
U Rotor Speed
W Relative Velocity
x Axial Position
Greek
Δ Difference
ΔP Change in Pressure
η Efficiency
𝜏 Torque
ΔT Change in Temperature
Subscripts
a Ambient
0 Total Property
1 Compressor Inlet
2 Compressor Exit
3 Combustor Exit
4 Turbine Inlet
5 Turbine Exit
Ave Average
h Hub
i Intake
j Jet pipe
m Mean
T Tip
ts Total-static
tt Total-total
Abbreviations
1 Introduction
1.1 Background
An axial flow turbine needs to be designed. The turbine is to drive a mixed flow compressor in a micro
turbojet engine. The compressor design in the test engine has a pressure ratio of 3:1 at 100 000 RPM
however this speed was not reached during engine testing, the engine manufacturer has advised that the
design point speed is 95 000 RPM and that 100 000 RPM was possible. The engine used for testing is a 180
N static thrust class micro turbojet, built for radio controlled (RC) model aircraft. This engine has an Inconel
600 alloy axial flow turbine rotor. The maximum engine speed is 100 000 RPM, while the design point speed
is 95 000 RPM. The turbine rotor diameter is 70 mm. The engine was tested and the results used to validate
Due to the size of the turbine rotor and the limited blade thickness of only 3 mm at the blade tips and 4,2 mm
at the hub, it was not possible to accommodate internal cooling passages due to high cost and complexity. It
therefore was not practical for this application. This is inherently why turbine inlet temperatures in micro
turbojets are significantly lower than those in large commercial or industrial turbojets. High by-pass ratio
turbofan engines have turbine stage inlet temperatures approaching 1700 °C, which is achieved due to
internal blade cooling. Compressed air is bled from the compressor casing and directed past the rotor shaft
and hub into the blades where it re-joins the flow having passed through the combustion process. Increasing
turbine inlet temperature is well established as a means of enhancing cycle efficiency for a turbojet cycle
engine. The proposed design would operate at a turbine inlet temperature (TIT) of approximately 600 °C.
Further research towards achieving an operating temperature of 850 °C with light weight titanium aluminide
Other performance improvements may be obtained by reducing the mass of the rotating components. This
lowers inertia in the assembly and would allow benefits such as fuel consumption and faster spool response
times. Commercial benefits of such technologies could result in smaller more powerful engines, reducing
drag, reducing mass and the engine reducing SFC (Specific Fuel Consumption).
1.2 Aim
The aim of this work was to design a single stage axial flow turbine rotor for a small gas turbine engine. In
conducting this work, the second aim was to develop design requirements by experimentally testing a micro
turbojet engine. The results obtained would then be used to validate design analysis numerical models.
1.3 Objectives
1. Conduct detailed literature study on the current state of the art in micro turbine design analysis
2. To design an experimental program to test the performance of a bench mounted micro turbojet
engine
3. Conduct detailed bench experimental tests to get performance data from an existing micro turbojet
engine.
5. Validate the CFD analysis software using the experimentally obtained data
6. Use the validated CFD model to analyse a single stage axial flow turbine of a micro turbojet engine
1.4 Scope
A cycle analysis for the turbojet must be completed. The cycle conditions that will be evaluated for the design
Design point (95 000 RPM with turbine inlet at 600 °C (actual) and 850 °C (theoretical in software
only).
Off-design, full power (at 100 000 RPM) with turbine inlet temperature of 1000 °C (theoretical in
software only).
The micro turbojet engine at the University of Johannesburg would be operated at the design point. The
altitude of the rig at the University is 1750m above sea level meaning experimental results would need to be
1.5 Methodology
This work will be based on a variation of the research and development methodology outlined in [1]. This will
be the approach followed throughout the project with the exception that a prototype will not be manufactured
and tested due to time and resource limitations. Engine testing will be used to validate the cycle analysis
To verify that the design parameters used in the Concepts NREC turbomachinery design code in terms of
stage pressure and temperature are realistic, the TIT and turbine inlet pressure (TIP) will be measured on
the test bench. Turbine exit pressure (TEP) and temperature (TET) will also be measured. The temperature
and pressure drop across the turbine would be indicative of the work done by the turbine. Correlation
between these experimental results and the design output data would verify the design in Concepts NREC
design code as well as the Gasturb 12 cycle analysis code when considering realistic component losses due
to engine size. The development methodology outlined in [1] is depicted in Figure 1-1.This thesis will focus
on the design elements up to and including the mechanical design elements work block. This mechanical
design will be limited to the rotor disk/hub, stator and rotor in the Concepts NREC design code. Finite
element analysis (FEA) of the design will be evaluated with 2 proposed materials.
completed in Gasturb 12. Thermodynamic design point studies will be covered in Chapter 4 and the results
of experimental work in Chapter 3 will be used to validate the methods in Chapter 4 to predict the turbojet
cycle performance of the test rig. Aerodynamic design and analysis will be performed using Concepts NREC
Axial and AxCent design codes. A turbojet engine test rig at the University of Johannesburg, Mechanical
Engineering Science Department will be used to test a 180 N micro turbojet engine. Test data will be
compared to the theoretical cycle performance and the results discussed. The thesis will be limited to the
2 Literature Review
This will be followed by an introduction to gas turbine operation, axial flow turbine theory and gas turbine
metallurgy and basic manufacturing principles. Finally, an overview of the micro turbojet will be presented
Driven by the technology development race during the Second World War, efforts by Whittle in the United
Kingdom and von Ohain in Germany led to the development of the modern jet engine [1]. The first jet-
propelled aircraft was from Germany and flew on August 27, 1939 [1]. The Gloster-Whittle aircraft made its
first flight on May 15, 1941 [1]. These two developments were solely aimed at the high speed military aircraft
applications. The initial engines had very limited life, very high fuel consumption and extremely poor
reliability. A long road of development was needed before the technology became available for civilian
In general terms, aircraft with a cruise speed of ~800 km/h or lower employ a propeller (turboprop) to achieve
best propulsive efficiency. [1] Aircraft with this propulsion system began service in 1953 and is still used in
today’s light aircraft. When aircraft to operate in the high subsonic regime, higher velocity exhaust jets are
required. This concept was supplemented by the addition of the fan by-pass duct, i.e. the turbofan engine. [1]
This engine type is essentially a multi-spool turbojet engine where a low pressure spool drives a fan stage
rather than a propeller stage. There are single spool turbofans an example is the French Snecma M53. The
marked difference is that the fan moves an even greater mass flow when compared to a propeller and thus a
lower mean jet velocity providing greater propulsive efficiency over the turbojet alone and hence a quieter
engine [1].
The compressor is driven by shaft power obtained from the turbine. The idea is that high pressure, a result of
the compressor work, and high temperature gas, a result of combustion, is expanded by the turbine to
produce shaft work and is further expanded to atmospheric conditions by the exhaust nozzle and thrust is
The air-standard open Brayton cycle is the ideal-cycle that represents the simple turbojet. It employs the
internal combustion process and the relevant heat transfer processes. The T-s and p-V diagrams depicting
the ideal cycle are illustrated in Figure 2-1 and Figure 2-2 [9].
Figure 2-1: T-s Diagram for Brayton Cycle, courtesy NASA, [40]
Figure 2-2: p-V Diagram for Brayton Cycle, courtesy NASA [40]
The practical gas turbine cycle differs from the ideal cycle due to irreversibilities that occur in the inlet,
compressor, combustor, turbine and propulsive nozzle. A derivation of the thermal efficiency of this cycle and
its dependence on pressure ratio will follow. The subscripts in the derivation relate to Figure 2-1 and Figure
2-2. Cycle pressure ratio may only be increased to a point, this is because of flow effects in decelerating
flows, which results in significant losses in the compressor diffuser as well as in the impeller. It is for this
reason that designers often look to increasing turbine inlet temperature when diffuser performance becomes
limited [1].
𝐶𝑝 (𝑇8 − 𝑇0 ) 𝑇0 (𝑇𝑇80 − 1)
𝜂𝑡ℎ = 1 − =1−
𝐶𝑝 (𝑇4 − 𝑇3 ) 𝑇 (𝑇4 − 1)
3 𝑇3
the gas constant. Both Cp and R differ with the working fluid.
Now considering the p-V diagram, where, the isentropic compression and expansion process, the pressure
𝑃3 𝑇3 (𝑘⁄(𝑘−1)) 𝑃4 𝑇4 (𝑘⁄(𝑘−1))
=( ) = =( )
𝑃0 𝑇0 𝑃0 𝑇8
where k is a constant referred to as the ratio of specific heat capacities for the working fluid, k = 1.4 for air
Making the relevant substitutions from 2-2 into 2-1, then 2-1 may be written as follows:
1
𝜂𝑡ℎ = 1 − (𝑘−1⁄𝑘)
𝑃
( 3)
𝑃0
The equation 2-3 is valid, provided that the value of (k), remains constant. A characteristic of the Brayton
cycle is a very large amount of compressor work (referred to as back-work) in comparison to the turbine
work. The compressor and bearings may absorb as much as 80% of the work output of the turbine. [1] The
remaining 20% would be consumed by ancillary systems which take off the remaining mechanical power
(such as a turboshaft engine, which finds application in helicopters.). [1] In the case of the micro turbojet,
which has no other auxiliary shaft power removal, the compressor and bearings consume 100% of the
turbine shaft power for a given speed. If this power balance between the compressor and turbine was not
equal, the engine would continue to accelerate. The implication of such a relationship is that overall
efficiency drops rapidly with a decrease in compressor and turbine efficiencies. The remaining enthalpy of
Having mathematically described the thermal efficiency of the Brayton Cycle, a discussion of the practical
turbojet layout and configuration is useful. The turbojet engine produces thrust by means of a high speed
exhaust jet. This power is what remains after the compressor back work has been done. Typically the
turbojet has a single convergent nozzle downstream of the combustor and turbine rotor. For the standard
turbojet cycle, the stations through the engine are numbered as depicted in Figure 2-3 [1] [9].
Moving through the stage, air is inducted into the intake. The air mass moves onto the compressor where the
pressure rise occurs. This pressure rise is typically referred to as the compressor pressure ratio (CPR) and is
expressed as the rise of pressure from station 1 to 2. Compressors will almost always have a diffuser
section, not depicted in Figure 2-3. The diffuser decelerates the flow and recovers dynamic pressure.
Decelerating flows are complex and will result in irrecoverable losses hence careful consideration during
design is required for diffusers of both axial flow and radial flow compressors. [1]
The fluid, now compressed enters the combustor at station 3, where it is mixed with fuel and ignited in the
combustor. Due to flow effects in the combustor, pressure is lost when moving from station 2 to station 3.
Typical losses across the combustor range from a 3% decrease in well-designed combustors to as great a
drop as 8% in poorly performing combustors. [1] The working fluid is heated during combustion up to 1600 K
or greater depending on cycle design and combustor design. A portion of the compressed gas leaving the
compressor is diverted to cool down the hot gas leaving the primary combustion zone of the combustor. This
cooling of the gas is important because the temperature to which the combustion gas is cooled is the
temperature at which the turbine must operate. The temperature partially cooled gas, for the sake of
discussion, is now approximately 1150 K and this is the temperature of the gas at station 3 (Figure 2-3) .The
turbine extracts the energy from the gas as the gas passes through to station 4 (Figure 2-3), behind the
turbine. The gas entering the turbine will have passed through a nozzle guide vane (NGV) which conditions
the turbulent flow leaving the combustor so that it approaches the turbine at the correct angles and at greater
speed. In Figure 2-3 the inlet guide vane (IGV) is located at station 3.
The gas leaving station 4, now at a lower temperature and pressure than at station 3 often is conditioned
again in an exit guide vane (EGV) but this is not always the case. The hot gas then moves onto the
converging nozzle which accelerates the flow and further reduces its pressure to produce thrust. Where an
after burner is employed more fuel is added to the hot high speed gas and burned in the ‘end-gas’ after the
turbine rotor which still has sufficient oxygen for complete combustion. The gas then rapidly increases in
energy and proceeds through the convergent section of the nozzle but then continues to a divergent nozzle
which allows the flow to continue to pass an increasing mass flow despite being sonic. The net effect of the
afterburner is a significant increase in thrust with increased fuel consumption. This is not employed in
commercial turbojets with some exceptions such as the Rolls Royce Olympus engine that found service in
the Concorde aircraft. Military aircraft, which are transonic, almost always have an afterburner [1].
Maximizing the performance of the turbojet cycle is a function of increased CPR, mass flow rate and turbine
inlet temperature (TIT). The maximum allowable turbine inlet temperature is limited by material properties of
the turbine rotor and the NGV. Turbine inlet temperature is increased by increasing the fuel flow provided
combustion is complete (stoichiometric). This is often done in conjunction with an increase in CPR. The
effect of increasing CPR is evident when considering 2-3 for the Brayton Cycle. When considering a turbojet
design analysis, fixed polytropic efficiencies for the compressor and turbine are selected for evaluation
purposes. Specific thrust is an indication of engine efficiency, for instance, two identical engines with
different specific thrusts might exist. The engine with the higher specific thrust is more efficient as it produces
more thrust for the same mass flow rate. SFC is the amount of fuel an engine will consume and is an
indication of how compact an engine is. Low SFC implies greater efficiency from a comparatively smaller
The specific thrusts in Figure 2-4 are greater than would be realised in a micro turbojet, however, the
relationship between TIT and CPR along with their relative effects on SFC and specific thrust are clear.
The role of the turbine in the turbojet is to provide power to drive the compressor and associated
accessories. This is accomplished by expanding high temperature and pressure gas to a lower temperature
and pressure. The magnitude of the pressure and temperature drop across the turbine is indicative of turbine
power produced, provided this is done efficiently and that the drop in pressure and temperature is as a result
of conversion of energy in the hot gas to shaft work delivered by the turbine. Temperature drops may also be
attributed to heat transfer in the NGV, rotor and shaft. In the turbojet, as the combustion gases expand from
the combustion chamber and through the turbine nozzle guide vane, it is accelerated. Simultaneously, the
gas is imparted with a swirl or tangential component as it is directed to the turbine blades. As the gas passes
through the turbine blade passage, energy from the gas is converted to rotation in the turbine and the
available energy in the gas is reduced. The torque generated by the turbine is determined by the mass flow
rate of the gas and the exchange of momentum of the gas stream between the inlet and the outlet of the
turbine rotor as well as the rotor tip radius and hub-to-tip radius ratio [1], [5].
A greater expansion ratio through a turbine indicates that a greater amount of work extracted from the flow.
Expansion ratio across a turbine is similar to, yet the opposite of the compressor pressure ratio except that
instead of a pressure rise in a compressor requiring work, pressures drop across the turbine provides work.
Conventional turbine designs remove the swirl component with the EGV. The reason for doing this is to
minimise nozzle related problems such as jet pipe vibration. For the micro turbojet however, the nozzle is
short and rigid so EGV’s are not necessary and omission of the EGV also reduces cost [1], [18].
There are two basic types of turbine, the axial flow turbine, as in the KJ66 in Figure 2-11, and the radial
Gas turbines for propulsion mostly make use of the axial flow turbine. At low mass flow rates, the blades of
an axial flow turbine become very small. The effect of this is that the tip gap losses become significant and
difficult to maintain. Small non-propulsion turbo machines, such as the automotive turbocharger, make use of
small radial flow turbine rotors, the reason for this is that the specific speed of the radial flow turbine is lower
Radial turbines are best suited to single stage, low cost applications. They are less efficient than axial
turbines but are far cheaper to manufacture. A typical example of a single stage radial turbine of low cost is
found in the automotive turbocharger. It features a radial flow turbine coupled to a radial flow compressor by
a common shaft. The reason they are used in this application is that maintaining the blade profile accuracy
for small axial flow turbines is excessively expensive. The radial flow turbine is significantly less sensitive to
blade profile tolerances and requires fewer blades per stage. Multi-stage radial turbines are rare because the
flow path between stages is torturous in that the flow must traverse several radial to axial bends and thus
In Figure 2-6, the nozzle is used to accelerate the gas, the absolute velocity is represented by C2 which
leaves the nozzle at angle α2 relative to the inlet velocity. The rotor inlet tangential velocity is U2 and the
nozzle inlet gas velocity is represented by V2. The velocity triangle in Figure 2-6 at the rotor exit is at the
mean radius of the exit, this is sufficient for illustrative purposes. U3 is the rotor velocity at the mean radius.
The gas leaves the rotor with relative velocity V3 at angle β3. The magnitude of the absolute velocity is C3
In an axial flow turbine such as in Figure 2-7 the flow enters the stator/nozzle guide vane parallel to the axis
of rotation. The flow is then turned within the nozzle and accelerated to a greater tangential velocity. The
tangential component enables the rotor downstream to extract angular momentum from the flow thereby
producing torque and power. The absolute velocity leaving the nozzle is C2 at an angle α2, relative velocity is
V2 at an angle β2. At the turbine exit after expansion in the rotor-blade passages the gas leaves with relative
velocity V3 at angle β3. The magnitude and direction of the absolute velocity at exit from the rotor C3 at an
angle α3 are found by vector addition of U to the relative velocity V3 and α3 is known as the swirl angle
Figure 2-8: Axial flow turbomachinery, courtesy Rolls Royce online [41]
Turbomachines fall into 3 basic flow categories. Axial flow, radial flow or mixed flow. Figure 2-8 illustrates the
axial flow path of gas in a compressor and in a turbine. [3], [31]. For this thesis an axial flow turbine will be
Turbines are either impulse driven, reaction driven or some combination of the two types. Most turbines
behave as impulse-reaction turbines and are typically 50% impulse and 50% reaction. In impulse turbines
the total turbine stage pressure drop occurs across the nozzle guide vanes. In the guide vanes, which are
analogous to a convergent duct, the flow is accelerated and the pressure drops. The gas then contacts the
rotor blade and imparts an impulse force as the high speed gas hits the pressure surface of the rotor blade.
This is the principle of operation of the Pelton Wheel [1], [3], [31].
In the reaction turbine the fixed inlet guide vane alters the flow direction of the gas and the flow velocity. The
rotor blade passage is equivalent to a convergent duct and accelerates the gas and decreases its pressure.
The difference in pressure on the pressure and suction surface of the blade results in a reaction force and
momentum exchange forcing the rotor to turn. Impulse-reaction machines have reaction at the blade tips and
behave as impulse turbines at the blade hub, this is to make full use of the benefits of each degree of
Figure 2-10 Velocity profile through impulse and reaction steam turbines [31]
Micro turbojets are used as propulsion systems for unmanned aerial vehicles (UAV’s) and in some instances,
medium and long range cruise missiles and guided bombs. Cruise missiles use larger turbojet engines but
they operate on a similar principle as the micro turbojet engine. These engines are simple turbojets typically
with mass flow rates ranging from 0,1 kg/s to 1 kg/s depending on performance requirements and
dimensional constraints. The compressor for a typical design would be a radial flow compressor connected
by a shaft to an axial flow turbine rotor. The compressor would have a radial-to-axial cross-over diffuser
feeding into an annular combustor located between the compressor and turbine, concentric to the shaft.
The size of a micro turbojet is such as that in Figure 2-11 sets it apart from typical turbojets. Micro turbojets
have radial flow compressors, almost all other turbojets are multi-staged axial flow machines. Manufacturing
tolerances for the micro turbojet is of an industrial scale to minimise manufacturing costs. One of the
implications of this is that the compressor and turbine blade tip clearance has a significant effect on leakage
and subsequent loss in engine performance. Larger engines suffer less from associated tip clearance and
secondary losses because a 1mm tip clearance on a 300 mm long turbine blade has a minimal leakage
mass flow compared to the bulk flow. However, on a micro turbojet such as the KJ66 which has turbine
blades that are 9 mm in height and may have a tip clearance of 1 mm, 11% of the blade height, significant
mass flow may pass through this tip clearance and therefore cannot be used to generate turbine shaft work.
The tip gap losses are associated with low Reynolds numbers which are characteristic of small jet engines.
This affects the thrust by creating turbulent vortices in the exhaust stream and boundary layer effects in the
Another implication of the small engine size is that there are no internal blade cooling passages. The turbine
must then operate at significantly lower temperatures, typically 760 K – 900 K. Large commercial jet engines
have turbines operating at 1950 K with sophisticated internal cooling passages such as the blade shown in
Figure 2-12. To increase the operating temperature without introducing cooling passages advances in
materials such as titanium aluminides and ceramics are important. These materials could potentially be used
When considering turbine design, high turbine inlet temperatures are desirable but the result of this is that
special high temperature alloys are then required. These alloys are expensive both in terms of material and
manufacturing cost. In an uncooled turbine, as would be designed for in the micro turbojet, the turbine
material is often the cheapest commercially available and this limits the engine performance
clearance flows. This statement, though based on empirical correlations for large commercial and industrial
gas turbines, holds true for micro gas turbines, albeit, in a more severe manner [37].
Profile loss denoted by Yp in most of the literature, is defined as the loss in stagnation pressure across a
blade row, divided by the difference between stagnation pressure and static pressure at the blade outlet [37].
𝑃01 − 𝑃02
𝑌𝑝 =
𝑃02 − 𝑃2
result of incidence ratio as well as corrections for thickness-chord ratios, but how well these apply to the
losses in micro turbojets is not known. In the open domain, this information is not readily available because
micro gas turbine applications often do not warrant the extensive test and evaluation programs of larger
commercial and industrial gas turbines. Great care must be exercised when making assumptions that are
commonly made in large gas turbines, these do not always hold for small machines [37].
The next area associated with total pressure losses is secondary loss. These losses exist because of three-
dimensional flows that are established by the state of end-wall boundary layers. Losses in stagnation
pressure are a major fraction of the total pressure loss and this is a result of the radially inward convection of
the end wall boundary layers on the suction surface side of the blade. The secondary loss coefficient, Ys,
also found to be true for large gas turbines by Ainley is given by [37]:
𝐶𝐷𝑠 cos 2 𝛼2
𝑌𝑠 =
𝑠⁄𝑙 cos3 𝛼𝑚
(s) is the space between blades and (l) is the blade chord for an arbitrary blade. As is conventional α 2 is the
local gas outlet flow angle and αm is the gas outflow angle at the mean line. However work by Dunham and
Came [37], proved that this was not correct for small turbines with low aspect ratio blades and modified this
𝑙 cos 𝛼2
𝑌𝑠 = 0,0334 ( ) ( )
𝐻 cos 𝛼1
2-6 Dunham, Came, modification for secondary loss in small gas turbines
It is worth mentioning that this does not imply this equation is at all suitable for micro gas turbine engines
which have appreciably lower aspect ratio blades than even small gas turbine engines [37].
Finally, tip clearance losses may be represented by Yk, such that, [37]
𝑙 𝑘 0.78
𝑌𝑘 = 𝐵 ( ) ( ) 𝑍
𝐻 𝑙
shrouded blades B = 0.25. When applied to micro gas turbines, blade shrouding is not common for various
Data on mean Reynolds numbers was collected by Ainley and Mathieson (referenced in [37]) which was for
Reynolds numbers of around 200 000 based on mean chord and rotor passage flows and exit flow
conditions in large gas turbines. They then went on to recommend that for lower Reynolds numbers, in the
order of 50 000, as would be expected to be found in a micro gas turbine, a correction to the stage efficiency
must be made. The rule of thumb for this correction was the relationship defined in equation 2-8 [37]:
−𝟏⁄
(𝟏 − 𝜼𝒕𝒕 ) ∝ 𝑹𝒆 𝟓
⁄5
𝑌𝑝 + 𝑌𝑠 ∝ 𝑅𝑒−1
2-9 Dunham, Came, Reynolds number correction in terms of profile and secondary loss
It is these total pressure losses that must be carefully considered when performing cycle calculations with
assumed component efficiencies for micro gas turbines. Equations 2-8 and 2-9 are mutually exclusive
because 2-9 is optional where Yp and Ys may be substantial terms as is the case in micro turbomachinery
sensitive to losses. In larger machines, these terms are sufficiently small such that the option need not
always be taken.
Historically, early turbine blades made use of ferritic and austenitic steels but in modern turbines these steels
are replaced by nickel based super alloys. [1] By adding nickel to the chemical composition, fatigue life and
the fatigue resistance of the turbine is significantly improved. The addition of chromium enhanced corrosion
resistance. The appropriate balance of these alloying elements, developed specifically for gas turbines, is
called Inconel. This is a proprietary trade name for a particular composition. Inconel alloys are widely used in
military, commercial and industrial gas turbines. Key to this widespread use is extremely low coefficients of
thermal expansion within a temperature range of 500 °C - 760 °C. These alloys have exceptional tensile
strengths of over 800 MPa in the 500 °C - 760 °C temperature range [6].
Turbine blades must resist fatigue and thermal shock stresses associated with high temperature and high
frequency gas fluctuations during operation. These blades must further resist corrosion and oxidation
processes in the hot stream during operation. Some small turbine engines often used in helicopter
applications feature dual alloy turbine disks, where the disk material is different to the blade material allowing
Nozzle guide vanes (NGV’s) are stationary in the gas turbine and are not subjected to rotational stress.
Primary material selection for NGV’s are based on thermal loading. Nickel based alloys are often used but
even these alloys require some form of cooling. It is not uncommon to have NGV’s coated in a sacrificial
ceramic coatings where internal cooling is not an option. These coatings reduce cooling requirements or
Turbojet engines make use of nickel based super alloys. These alloys contain 12 or more alloying elements.
Nickel super alloys have unique thermal and mechanical properties. They have extremely high corrosion
resistance at elevated temperatures and have among the lowest coefficients of thermal expansion of any
material in the temperature range of 500 °C – 760 °C. Within this temperature range it is not uncommon to
find tensile strengths exceeding 800 MPa. Nickel super alloys are usually cast, this is advantageous when
considering the complexity of internally cooled turbine blades such as Figure 2-12 [3].
Turbine blades are typically cast alloys. There are other blade forming techniques such as powder sintering.
Powder sintered blades are approximately 10% stronger and allow higher speed operation due to finer
crystal structures and more, smaller grain boundaries, this finds application already in the GE90-115B turbo
fan core engine in the low pressure turbine stage as well as the Rolls Royce Trent 1000-D/E [3].
Microstructural analysis of the turbine blade material shows that conventional turbine blades are comprised
of a matrix of crystalline structures arranged in a random manner. It was soon realised that arranging these
crystals in a longitudinal manner from blade root to tip significantly improves the service life of the blade. The
method used in forming these columns of crystal is called, directional solidification, see Figure 2-13 and
Figure 2-14. This ‘tuning’ of the grain structure of the blade material was further extended by growing the
blades from a single crystal of material. This had a major effect on creep and fatigue life of turbine blades.
Intricate casting ability for turbine blades is advantageous in commercial and industrial turbomachinery
because of the complexity of internal cooling passages within the blade [3].
Figure 2-13: Time to fracture based on casting techniques for turbine blades [3]
Figure 2-13 gives a relative indicator of the effect of crystal structure on elongation and time to failure, this
Over the last 30 years research has been conducted worldwide with the goal of applying lighter, stronger
materials in gas turbine hot side components. This research has led designers to the class of materials
called structural ceramics. Ceramic monolith components in gas turbines, such as that in Figure 2-15 have
the capability of increasing thermal efficiency and power output whilst reducing emissions. The superior high
temperature strength, durability and other physical properties make silicon nitride (SN) a prime candidate in
the silicon based ceramic group. Potential and demonstrated methods of achieving improved fracture
toughness, strength and high temperature performance combined with high temperature environmental
Figure 2-15: Ceramic, axial flow turbine rotor, courtesy online resource Allied
The material properties that limit the use of silicon based ceramics include high temperature oxidation,
fracture toughness, creep and fatigue, environmental degradation and high temperature strength. By
definition, fracture toughness is a property which describes the ability of a material containing a crack, to
resist fracture. Creep is the tendency of a solid material to move slowly or deform permanently under the
influence of an applied stress. When turbine blades rotate, they creep outward due to centrifugal forces and
temperature induced changes in tensile strength acting on the blade. Fatigue resistance is the number of
stress cycles of a specific type, such as tension and compression that a specimen can withstand before
Ceramics are not common as a turbine material but they have been successfully tested is in radial turbine
rotors. Radial turbines have fewer, thicker more robust blades and they are generally smaller in size than
axial turbines. It is not just sufficient to have ceramic turbine, special attention must be paid to the combustor
liner and exhaust nozzle, which also carry a substantial thermal load [7].
material. These are simple castings with no internal cooling. This places a restriction on the micro turbojet
performance as TIT is limited for safety concerns. Ceramics have been employed but at great cost and are
generally not considered. Aerodynamic losses for large turbomachinery, particularly secondary losses, are
not easily quantified or documented in any literature for micro turbojets. Estimations may be presumed, but
there is no conclusive supporting documentation that suggests these losses extend further than Reynolds
number effects. These losses are however substantial and have a major influence on the micro turbojet
performance. Micro turbojets are not as efficient as larger scale gas turbines, these effects will be explored in
Johannesburg. Temperatures and pressures were measured across key components such as the inlet, the
compressor, combustor, turbine rotor and exhaust nozzle. There were various limitations to the experimental
programme that limited the extent of testing. These are expanded upon in the next section. The results were
3.2 Limitations
Ideally before testing, the engine should have been completely disassembled and the individual components
3D scanned and measured. The point being, the 3D model would be then used in a simulation of the as-built
components to get the most realistic simulations of the engine performance. This will facilitate a direct
comparison within the Concepts NREC design code and Pushbutton CFD for the proposed design. This
detailed model would have allowed critical measurements to be made in the area of the turbine and NGV to
ensure the proposed design has the same running clearances, rotor shaft diameter and exhaust cone
Due to restrictions at the University, the test engine was not disassembled. However, the engine
manufacturer did provide an NGV and a turbine rotor to be scanned. It could not be confirmed that these
were the exact components in the test engine, measurement of critical dimensions would have confirmed
this but this was not possible due to disassembly not being done. The scanning of these individual
components was very poor, part of the problem being the very small flow passages between the NGV vanes
and rotor blades. This caused reflections of the laser even when stained white with NDT dye developer and
scanned with a Faro laser arm. The cost of the scanning was a limitation because of the length of time that
would have been required to get a high resolution point cloud suitable to fitting a surface model.
numerical models to be used in the design. The test engine was operated as close to 100 % (100 000 RPM)
speed as possible and data collected on critical performance parameters. This experimentally obtained data
was used to validate the cycle evaluation software code Gasturb 12.
Tests were conducted for one minute at maximum attainable engine speed. Two tests were conducted per
day of testing, a total of twelve test days were completed. Testing was conducted in the late-morning and
mid-afternoon on alternate days. Testing was not conducted during rainy weather or when humidity was
greater than 50 %. The data was reduced to one decimal and time averaged across all test days. There were
3 data runs per test. Two further tests were conducted at 92 % and 95 % spool speed.
The purpose of the experimental testing of a micro turbojet was to collect data for a working engine with
Inconel 600 alloy turbine rotor and radial out-flow compressor at design point. Provided the test results are
within an acceptable range of the Gasturb simulations, the result of the cycle analysis could be assessed as
acceptable. This potentially acceptable cycle validation of Gasturb and user specified component efficiencies
would form the basis for the design of the rotor in the concepts NREC design code.
The parameters that were measured included pressure, temperature, fuel flow rate, thrust and engine RPM.
Ambient pressure was measured using a barometer while the compressor and turbine pressures were
measured using a custom made u-tube manometer made and calibrated at the University. Transducer
Combustor pressure drop measure with U-tube manometer (0 kPa – 350 kPa) 1 % resolution
For the compressor exit pressure the tap was located after the radial diffuser in front of the combustor inlet.
No internal photographs were taken as the University prohibited disassembly of the engine for this purpose.
Engine speed, Hall effect sensor (0 RPM – 130 000 RPM) 1 % resolution
The thrust measurement load cell was calibrated by means of applying a known load to the load cell and
measuring the transducers response. The cell was loaded with a 5 kg mass, 10 kg mass and 15 kg mass.
The transducer characteristic is linear up to 50 kg, linearization is used to interpolate the thrust or applied
load.
The pressure transducers were calibrated with a hand pump owned and operated by JAD Systems. JADS
supplies transducers and data acquisitions systems by National Instruments to The University of
Johannesburg. Thermocouples were calibrated using precision temperature controlled oil baths at the CSIR
Figure 3-1 presents the static thrust results measured on the test rig.
80 700
74
70 574 589 588 600
566
530 538 63.2
50
44.9 400
40
36.7
300
30
25.7
Thrust 200
20
TIT (Celcius)
12.4 100
10
6.1
0 0 0
0 10000 20000 30000 40000 50000 60000 70000 80000 90000 100000
Engine RPM
In Figure 3-1 the test engine at the University of Johannesburg produced 74 N static thrust at 91 740 RPM.
Full engine speed was not achieved during testing. Possible reasons for this will be discussed in Chapter 4.
Figure 3-2 presents the variation of compressor pressure ratio and pressure with engine thrust.
2.5 250
2.25
Compressor Pressure Ratio (CPR) 1.98
2 1.88 200
1.79 199.4
1.67 175.4
1.55 166.6
158.6
Pressure (kPa)
1.5 148.0 150
1.27 137.3
1.12 112.5
1 99.2 100
88.6
CPR
0.5 50
Pressure kPa
0 0 0
0 10 20 30 40 50 60 70 80
Static Thrust (N)
The maximum pressure ratio presented in Figure 3-2 (CPR) calculated from measured data is 2,25:1 and
was measured when thrust was 74 N. As expected the maximum thrust measured was at the highest
pressure ratio of the compressor, but this was not at design speed.
Figure 3-3 presents the variation of mass flow rate and engine thrust as a function of the engine rotational
speed.
0.35 80
74
0.3 70
63.2 0.287
57 60
0.25 0.265
Mass flow rate (kg/s)
0.221 0.243
The mass flow rate shown in Figure 3-3 increases with engine speed. This is expected because for a
subsonic compressor at constant diameter the mass flow rate is proportional to the RPM, tip diameter D and
air density ρ, 𝑚̇ ∝ 𝜌𝑁𝐷. Pressure ratio is proportional to air density, ρ and the square of RPM and diameter
mass flow rate of 0.325 kg/s is expected. The engine was unable to reach design speed of 95 000 RPM. The
maximum speed achieved during testing was 91 740 RPM with mass flow rate measured at 0.287 kg/s.
250
199.4
200
175.4 190.1
158.6 166.6
166.3
148.0 158.4
Pressure (kPa)
0
0 20000 40000 60000 80000 100000
Engine RPM
The measured combustor pressure drop in Figure 3-4 is between 5,4 and 5,5 % on the test rig. This
pressure drop is large. Typical pressure drops are around 3 %. However, this is on large combustors in
commercial engines. The hobby grade engine in the test rig is not designed for optimum combustion during
idle. Tolerances and clearances are of a much bigger percentage of the machine dimensions in the case of
the small engine. A further implication of the machine size is that the hole configuration/hole pattern of the
cycle code analysis done in Chapter 4 particularly in light of the manufacturers claimed thrust and expected
component efficiencies. Trends in the data are consistent with those expected of the typical turbojet cycle in
terms of mass flow rate, pressure ratio and engine speed. The trend in thrust in particular will be shown in
chapter 4 to have a reasonable correlation to the Gasturb 12 data. It is up to the user to specify realistic
component efficiencies and losses within Gasturb. These will be discussed in Chapter 4. Combustor
pressure drop was high at around 5,5 %, but this is expected for a very small engine. Typical combustor
pressure drop is in the region of 3 % for large engines. Tip clearance was measured at approximately 1 mm
with a Vernier calliper, considering a blade in this rotor is only 14 mm long, significant losses will occur with
the clearance being more than 7 % of the blade height. These losses are associated with tip clearance and
secondary flow losses due to the low aspect ratio (AR) of the blades. For the test engine turbine blade AR is
4 Cycle Analysis
parameters and component efficiencies in a turbojet engine or other gas turbine. These cycle calculations
give an indication of overall performance for, in this instance, the turbojet cycle. There are a great many
variables that make it impractical to have set algebraic expressions for the efficiency of a real cycle. The
method and principles outlined in this section can be readily used to evaluate the performance of a turbojet
4.2 Aim
For this thesis, the operating points to be evaluated in the cycle code are sea-level static (SLS) conditions
and altitude static (AS) at 1750 m above sea level. The aim is to attempt to calculate, as accurately as
possible, the test rig performance within the Gasturb 12 cycle evaluation software code. Good correlation
with the experimental data would result in a fair degree of confidence in the cycle evaluation performed in
Gasturb 12 for the proposed turbine rotor design. It should be mentioned that Gasturb 12 makes use of
empirical correlations and methods for large commercial gas turbines. It should then be clear that a high
degree of correlation with the performance of micro turbojets is not likely because of the nature of the loss
mechanisms and loss effects experienced by micro turbojets due to their small size.
Using the Gasturb 12 cycle code the variables that control thrust are:
Compressor efficiency
Turbine efficiency
Auxiliary component losses, such as bearing losses, combustor losses and altitude effects
Fixing compressor efficiency, turbine efficiency, power off-take (which is zero for this engine) and
compressor pressure ratio, the thrust can be increased by either raising the mass flow rate or by increasing
the turbine inlet temperature. Turbine inlet temperature has an upper limit of approximately 600 °C (873,15
K) in an uncooled turbine rotor cast from Inconel 600. It is a conservative temperature limit imposed due to
safety of operation. This TIT of 873,15 K is used to quantify the turbine inlet temperature upper limit at
design point for the turbine rotor design. The 873,15K TIT can be exceeded for short periods of time
anywhere within the flight envelope, according to the engine manufacturer. Exactly how long the TIT may be
exceeded is not disclosed by the manufacturer. In Gasturb 12, the stations are numbered according to
Figure 4-1. For illustrative purposes, an axial flow compressor is shown. Other capabilities of the cycle code
are depicted in blue text in Figure 4-1. These were not used for the cycle evaluation.
The cycle specifications for the basic cycle analysis are listed as follows, but corrections were made to
CPR 3,1:1
not being achieved. According to the engine manufacturer the maximum engine speed is 100 000 RPM, with
cruise speed being 95 000 RPM. There was no technical specification sheet available for the engine.
However, using Gasturb 12 cycle code and analysing the testing cycle results for the engine with the
from measurement)
80 % isentropic efficiency for the turbine (estimation based on losses in micro turbojets, [34] )
To obtain a useful cycle analysis for the experimental work a complete discussion on the individual
component efficiencies is required so that these may be used with some confidence. Firstly, from the data
obtained experimentally, the actual isentropic efficiency of the compressor and the turbine may be calculated
with a fair level of accuracy. The combustor efficiency couldn’t be evaluated as it is an extremely difficult task
to complete experimentally without a dedicated combustion rig where gas composition could be analysed
thoroughly. In general, conservative practice on large gas turbine combustors is to assume 98 % isentropic
efficiency for the purpose of cycle calculations [37]. On micro gas turbines, it is not likely to be much better
than this, more likely worse. However there is no literature or data base on combustion efficiency for micro
turbojet engines. For this reason, 98 % isentropic combustion efficiency [13] will be used for the cycle
analysis of the experimental work and for all subsequent cycle analyses from this point on. Combustor
pressure drop however was measured at 5,5 % and this will be used.
The isentropic efficiency of the compressor can readily be calculated using equation 4-3. The result is that
with the compressor exit temperature of 410 K and inlet temperature of 294 K, the isentropic efficiency of the
compressor on the test rig at test conditions is approximately 68 %. At design point the manufacturer claims
75 % but this could not be verified as the test engine did not reach the design point speed. Turbine isentropic
However, the reason for not calculating this value using equation 4-7 is that the isentropic efficiency is
calculated based on stagnation temperature which was not measured during testing. To test the validity of
the turbine isentropic efficiency of 80 %, a parametric study of the turbine efficiency is plotted as a function of
the net thrust measured during testing with the other experimental data. The range of isentropic turbine
efficiencies tested is 20 % to 90 %. The thrust measured during testing was 74 N. Therefore, according to
the parametric plot, the turbine isentropic efficiency would be in the region of 79,5 % in Figure 4-2, and thus
the 80 % estimation from [34] is representative for the purpose of cycle analysis in this thesis.
Figure 4-2: Parametric study of the isentropic efficiency for the turbine
Based on specific speed and specific diameter relationships as discussed [13], estimates for the design point
isentropic efficiencies of the radial compressor and axial flow turbine were estimated as 68 % and 80 %
respectively. Figure 4-2 was created in Gasturb 12 using the cycle parameters used in the cycle analysis.
The mechanical transmission loss due to bearing windage is not known for the test engine. Due to
restrictions at the University, a complete disassembly of the test engine was not possible. As a result,
bearing identification was not possible and hence bearing loss calculation was not possible. However,
according to [39], bearing loss and compressor back face windage loss can account for 5 % loss in power
transmitted between the compressor and turbine. Once more this is for larger engines, no reliable data is
available for micro turbojet engine bearing systems, hence a mechanical efficiency of 95 % will be used.
4.4 Methodology
The cycle conditions that will be evaluated for the design and the test rig will be:
Design point, manufacturers claimed data ( 95 000 RPM with turbine inlet at 873,15 K (actual) and
1123,15 K (theoretical) )
Off-design, full power, manufacturers claimed data ( 100 000 RPM with turbine inlet temperature of
1273,15 K (theoretical) )
Component efficiencies used were those discussed in section 4.3. In this section, experimental data cycle
analysis will be referred to as condition 1. Design point at manufacturer’s claims, as condition 2 and off-
design as condition 3. Condition 2 and condition 3 have theoretical TIT values for operation of 1123.15 K and
analysis covering a suitable range of the turbine inlet temperature and the compressor pressure ratio was
conducted. The cycle corrected for SLS conditions as defined since the basic cycle described in section 4.2,
compressor outlet pressure. The other parameters in the cycle were not changed. The range of pressure
ratios investigated was from CPR 1:1 to CPR 4,9:1. The basic cycle design point, see Figure 4-9, is indicated
cycle net thrust range for the given pressure ratio, this appears to be between 3:1 and 3,5:1. Increasing the
pressure ratio further than this range would actually reduce the thrust. Considering that a pressure ratio of
4,7:1 would result in 100 N of thrust. The basic design cycle result at a pressure ratio of 3,1:1 depicted by the
dark point on the curve in Figure 4-3 shows a thrust of 110 N. It is evident that increasing pressure ratio on a
moderate efficiency compressor does not produce greater thrust. Had the compressor been more efficient
the optimum pressure ratio range would have been higher than that depicted in Figure 4-3.
Figure 4-4: Pressure ratio sensitivity analysis with 85% isentropic efficiency
Having made such a statement, the same pressure ratio range was evaluated but this time with a
compressor isentropic efficiency of 85%, the result in Figure 4-4 clearly indicates that the optimal compressor
pressure ratio range is farther up the pressure scale between 3.8:1 and 4.2:1 and can extend as far up as
4.9:1. Also clear is the dark point depicting SLS conditions on the curve, now lower than this optimal range
but clearly showing an increase in the net thrust produced due to greater compressor efficiency.
equivalent to the TIT. Furthermore, for the sake of simplicity, we expect temperature drop across the NGV to
be as low as possible) was varied and the effect on thrust plotted as a function of net thrust. The isentropic
efficiency of the turbine was fixed at 80 % in line with the estimations made earlier in section 4.3. The TIT
range evaluated was between 700 K and 1420 K and the results are plotted in Figure 4-5 and Figure 4-6.
The dark point on the curve in Figure 4-5 is the SLS cycle condition described in section 4.2. It is clear from
the data presented in Figure 4-5 that increasing TIT increases net thrust. There is an almost linear
relationship between 900 K and 1420 K that suggests every increase of TIT by 10 K increases net thrust by
2,61 %.
Considering this information, restraint should be exercised when chasing greater net thrust from a cycle
because increasing TIT to gain thrust has severe implications in the turbine design and material selection as
well as on fuel consumption. It is most likely more beneficial to seek greater turbine efficiency with only a
moderate increase in TIT. Having said this, an investigation into the effect of turbine efficiency on net thrust
is worth investigation.
that typical axial flow turbine isentropic efficiencies for large mass flow rate turbines is in the range of 85 % to
as high as 90 % for some steam turbines. This would not hold true for small axial flow machines with
relatively large tip clearances, low aspect ratio blades and comparatively small Reynolds numbers, but aids
in illustrating the effect of an increase of efficiency by an arbitrary 10 % over the design point estimation of
80 % and the effect on net engine thrust. The effect on net thrust is an increase to 126 N for SLS conditions.
That equates to an increase in net thrust of 12,7 % for an increase in efficiency of 10 % whilst not consuming
more fuel due to increasing TIT. With a lower efficiency shown in Figure 4-5, the TIT would have to be closer
to 950 K, an increase of approximately 8 %. This will have significant implications on turbine material and
rotor life for uncooled blades in micro turbojet application. It is for this reason that maximum isentropic
efficiency for the proposed turbine design should be of prime importance. When integrating the turbine
design into an engine, the compressor isentropic efficiency will need to follow peak efficiency as good gains
in net thrust can be found, to a point, without increasing TIT, which should be a last resort in the pursuit of
engine performance.
In relation to gas turbine thermodynamics temperature increase at the turbine inlet is desirable to a limited
extent such that material temperature limits are not exceeded. In terms of wanting to minimize engine size
for a given specific thrust, increasing turbine inlet temperature may be a technique employed. One may
readily calculate the actual thrust by multiplying the specific thrust by the mass flow rate through the engine.
It must be noted further that at a constant pressure ratio, increasing T03 will raise the specific fuel
consumption (SFC) [1]. The result of raising the pressure ratio is then a means of reducing the SFC. With T03
(turbine inlet temperature) held at a fixed value, raising the pressure ratio initially results in an increase in
thrust but eventually leads to a decrease in thrust as the turbine is unable to supply enough power to drive
the compressor at overly high design pressure ratios. It should be apparent that the optimum pressure ratio
increases for maximum thrust/power as the value of T03 increases. This theory holds true for engines
operating in a subsonic environment and stationary engines. Furthermore, during the calculation process the
turbine expansion ratio is determined and the turbine discharge and jet nozzle are evaluated for choking
conditions. Parameters such as specific fuel consumption (SFC), net thrust (F), fuel flow (mf), turbine
discharge nozzle area (A5) and the discharge velocity of the jet stream (C5) are also determined [1].
The method described in this section and the equations used in this section are from [1]. The Gasturb 12
code uses a similar set of equations with the exception that certain parameters are modified to capture real
effects from real experimental data on various different gas turbine configurations. For the calculations that
follow in this section, the station label convention is presented in Figure 2-3. The variables in this section are
Px which is the pressure at arbitrary location X and P0x is the stagnation pressure at X. Similarly, Tx is the
static temperature at arbitrary location X and T0x is the stagnation temperature at X. Ca is the local speed of
sound, Cp is the gas coefficient for the working fluid. The ratio of specific heat 𝛾 of the fluid is approximately
1,4 and 1,3 for air and combustion gas respectively. To properly define the stagnation conditions at the
intake it is necessary to accommodate the eventual performance at altitude and with a significant forward
speed component in addition to the stationary performance at sea level conditions. To allow for this the
following equation is used to quantify the ram pressure effect at the intake as a result of forward speed [1].
𝛾⁄(𝛾−1)
𝑃01 𝐶𝑎2
= [1 + 𝑛𝑖 ]
𝑃𝑎 2𝐶𝑝 𝑇𝑎
The temperature effects of the forward speed are considered in the following equation.
𝑐𝑎2
𝑇01 = 𝑇𝑎 +
2𝑐𝑝
For the analysis the jet engine was considered in stationary form at sea level conditions, the effect of this is
the term Ca reduces to zero because it is a product of the forward flight speed Mach number and the local
speed of sound. When this is the case T01 =Ta and equation 4-1 is equal to one. Next we may begin to
handle the conditions across the compressor, from stage 1 to stage 2. This is done in the following manner
[1].
𝑃02
The pressure ratio across the compressor is described by the ratio ( ). Since all variables are known
𝑃01
except𝑇02 , this can be determined. It should be pointed out that 𝑃02 is determined by knowing the pressure
ratio and ambient pressure. Proceeding to the analysis from stage 3 to staged 4 with the temperature 𝑇03
From this 𝑇04 is determined. This is the actual temperature of the gas leaving the turbine toward the nozzle.
Knowing the combustor pressure drop Δ𝑃𝑏 the pressure leaving the combustor may be calculated using the
Δ𝑃𝑏
𝑃03 = 𝑃02 (1 − )
𝑃02
To determine the actual pressure of the gas leaving the turbine it is necessary to first determine the ideal
′
temperature 𝑇04 of the gas leaving the turbine. This is done by making an approximation of the turbine
efficiency 𝜂𝑡 at the analysis point. The equation that allows this to be done is presented below [1].
′
1
𝑇04 = 𝑇03 − (𝑇 − 𝑇04 )
𝜂𝑡 03
′
With all the terms on the right being known 𝑇04 is determined by making the substitution of this result into the
following:
′ 𝛾⁄(𝛾−1)
𝑇04
𝑃04 = 𝑃03 ( )
𝑇03
When the analysis takes into account the individual component losses the overall efficiency of the cycle
becomes dependent on both the pressure ratio and the maximum cycle temperature T03. For each maximum
temperature and a particular pressure ratio, the cycle thermodynamic efficiency has a maximum value.
Eventually the thermodynamic efficiency will fall for a continuously increasing pressure ratio because of a
necessary reduction in fuel supply to give the required turbine inlet temperature T03. This is a result of higher
compressor delivery temperature and is outweighed by the work required to drive the compressor. It is
apparent at the outset that low power engines require lower pressure ratios due to blade size limitations and
lower turbine inlet temperature because of the associated difficulties with cooling excessively small turbine
blades [1].
The actual turbine discharge pressure may be determined. This pressure 𝑃04 is pivotal in determining the
nozzle choking conditions which is evaluated by means of the nozzle pressure ratio, given by the following
expression:
𝑃04
𝑃𝑎
𝑃04 1
=
𝑃𝑐 1 𝛾 − 1 𝛾⁄(𝛾−1)
[1 − ( )]
𝜂𝑗 𝛾 + 1
We may then proceed to analyse the inequality that arises from comparing the nozzle pressure ratio to the
𝑃04 𝑃04
critical pressure ratio of Equation 4-8 with the result that where < we may say that the nozzle is not
𝑃𝑎 𝑃𝑐
𝑃04 𝑃04
choked and when > the nozzle is operating in the choked regime [1].
𝑃𝑎 𝑃𝑐
For the turbojet engine the choked condition at the nozzle (station 5 in Figure 2-3) is desired because the
useful thrust of this type of engine is largely created from the sonic flow stream leaving the nozzle. The sonic
conditions at the outlet require that the nozzles critical conditions 𝑇𝑐 and 𝑃𝑐 be determined using the
following. When the passage is choked, T5 is at critical temperature such that [1]:
2
𝑇5 = 𝑇𝑐 = ( )
𝛾+1
1
𝑃5 = 𝑃𝑐 = 𝑃04 ( )
𝑃04 ⁄𝑃𝑐
To determine the required nozzle area it is necessary to calculate the density of the gas at the nozzle and to
𝑃𝑐
𝜌5 =
𝑅𝑇𝑐
𝐶5 = √(𝛾𝑅𝑇𝑐 )
The nozzle area required to satisfy the sonic condition is then determined by inserting the results of Equation
4-11 and 4-12 into the following, provided that the mass flow rate has been selected prior [1]:
𝑚̇
𝐴5 =
𝜌5 𝐶5
With the relevant nozzle area determined the theoretical specific thrust Fs may be calculated [1].
𝐴5
𝐹𝑠 = (𝐶5 − 𝐶𝑎 ) + (𝑃 − 𝑃𝑎 )
𝑚̇ 𝑐
In this section the results of the cycle analysis using the Gasturb cycle code for the experimental data
collected at 91 740 RPM, manufacturer’s claimed design point speed of 95 000 RPM and finally at the off-
design point 100 000 RPM will be presented and briefly discussed. Component efficiencies will be fixed at
the manufacturer’s claimed values for the cycle at 95 000 RPM and at 100 000 RPM. Mechanical loss values
will not be changed from those rationalized earlier. Similarly combustion efficiencies will be unchanged at
lower than typical values due to micro turbojet limitations as discussed earlier. Station numbers of previous
sections do not apply as this section numbering convention is dictated by Gasturb in accordance with Figure
4-1.
obtained results.
The theoretical thrust as determined in Gasturb is approximately 64.6 N at 1750 m altitude. The measured
thrust was 74 N at 1750 m from section 3.7 with TIT = 861 K. This moderately close (13 %) correlation to the
test data for thrust suggests the Gasturb results for thrust and component efficiencies are realistic at the
Figure 4-8: SLS corrected Gasturb approximation of experimental results using test
data as input variables
SLS corrected data is shown in Figure 4-10 from the Gasturb code, component efficiencies were not
changed, ambient pressure, temperature and air density, thrust is 71,4 N at SLS. Looking at the data at SLS
conditions, it was expected that the thrust would increase as the intake air pressure is greater. Correcting for
altitude shows the mass flow rate would be higher than measured at AS, this is consistent with air density at
At the engine manufacturer’s recommended operating temperature of the test rig of 600 °C (873.15 K), cycle
analysis yields a theoretical thrust value of 109 N at SLS conditions. The cycle summary is presented in
Figure 4-10. At this TIT, the engine theoretical thrust is significantly lower than expected. 180N of thrust was
expected as per the manufacturer’s claim but 109 N was achieved. This is a large discrepancy which can be
attributed to either a deficiency in mass flow rate, TIT or the manufacturer’s claim being incorrect. To obtain a
thrust of 180 N, with a quick calculation using Gasturb, with no change to the TIT or component efficiencies,
would require a mass flow rate of 0.55 kg/s. This is 41 % more mass flow than the manufacturer’s claimed
flow rate of 0.325 kg/s. Having discovered this, it is more likely that the thrust class of this engine supplied for
testing is closer to the 120 N engines which are also manufactured by the same company and not the 180 N
engines.
If this were the case, which it appears it very well is, the engine cycle analysis summarized in Figure 4-10 is
within 9,2 % of 120 N and is possibly within a 10 % performance band that small engine manufacturers are
likely to find acceptable given manufacturing tolerances and as-built characteristics of each component in
each engine being highly variable. An alternative approach to achieving the 180 N claimed thrust would be to
increase the TIT. However, being a small gas turbine, cooling of the rotor blades is not practical and hence
not possible. Another quick calculation with mass flow rate fixed at 0.325 kg/s and all component efficiencies
fixed, shows that a TIT of at least 1200 K would result in 180 N of thrust being produced by the engine.
Component life would be significantly lowered at this temperature and as such we can neglect this method
for achieving the 180 N thrust level claimed by the engine manufacturer.
Figure 4-10: Cycle result file for condition 2 at SLS and 873.15K
When the cycle is evaluated at the theoretical temperature of 1123.15 K, the thrust is significantly increased
to 170N. No other parameters were altered. It is clear increased TIT shows an increase in thrust for the
Successfully implementing 1123.15 K as a revised design point TIT shows the potential benefit of operating
at greater turbine inlet temperatures even though operating life would be lower. A complete study into
material properties of an uncooled alloy in a gas turbine environment would need to be completed to fully
understand material behaviour. Various tests on material properties would have to be investigated from
tensile testing, to corrosion testing and heat treatment methods to address shortcomings. Long term
exposure characteristics, such as creep and high cycle fatigue are also important to understand. However, in
Figure 4-13 and 4-14 present the analysis results for condition 3 i.e. at TIT of 1273,15 K at sea level.
Figure 4-13 Station properties for condition 3 off-design at SLS and 1273.15K
Figure 4-14 Cycle result for condition 3 off-design at SLS and 1273.15K
Operation of the cycle with TIT of 1273,15 K (1000 °C), see Figure 4-14, results in a thrust of 204 N. This
operating cycle would be indicative of the off-design mode possible if a turbine could be designed to operate
at such an elevated temperature, with no internal blade cooling. With Inconel 600, this would not be possible
when considering that the temperature of the exhaust jet would be around 1000 K not to mention the NGV at
In general turbine jet engine testing and performance analysis is conducted at SLS ambient conditions where
the thrust is greatest. As the altitude of the engine increases, thrust decreases. This is due to lower intake
pressures which are associated with lower ambient air density. Ordinarily, lowering intake temperature would
increase overall cycle efficiency because the intake air density increases. The decrease in thrust is due to a
decrease in ambient pressure and more than outweighs the increase as a result of lower temperatures. The
net effect of this is that the thrust of any turbojet engine decreases with an increase in altitude. The mass
flow rate through the engine decreases and this affects fuel flow rate which will also decrease with altitude
Making reference to data in Figure 4-10 and Figure 4-12 increasing TIT from 873.15 K to 1123.15 K and
fixing all other parameters results in increased thrust from approximately 109 N to 170 N, this is expected.
For the gas turbine, the cycle efficiency increases with TIT. The only instance where this does not apply is
when turbine blade cooling becomes necessary and as the losses that follow are introduced. It is evident that
for a fixed CPR, it is desirable to increase TIT as much as possible to the safe material life limit. For the
design, 1123.15 K (850 °C) is more representative of the target objective TIT for future work with an
In many instances, the design point is not the maximum performance point for a particular engine. Some
engines are designed to have maximum fuel efficiency during cruise, and this is a stark contrast to the
maximum performance that the engine is capable of. With this in consideration, an off-design mode for the
cycle was evaluated at 1273.15 K (1000 °C), see Figure 4-13 and Figure 4-14. It should be evident then that
the performance gain is substantial when considering data in Figure 4-10 in relation to Figure 4-14, a 31.42%
increase in TIT from 873.15 K to 1273.15 K has resulted in a thrust increase from 109 N to 200 N without
These results must be tempered however. The problem with having high turbine inlet temperatures extends
beyond the turbine design itself. It has effects on the life of the combustion chamber and flame tube.
Specialized materials would have to be employed in the combustor to cope with the higher temperature of
the primary zone kernel and the cooling of the flame tube in the combustor. The effect of sufficiently
increasing turbine inlet temperature means a reduction in the air/fuel ratio. This implies less of the air is
The latest developments in heat transfer that may be worth considering include transpiration cooling. These
are cooling flows in a network of narrow passages within the flame tube wall which can result is a 50 %
reduction in flow through the combustor core flow. This could negate some of the gains of higher TIT by
reducing the effective volume flow rate as some of the gas is not expanded in the combustor by addition of
fuel and combustion. Whether such a cooling method could be employed in a micro turbojet engine is
dependent on the cost of the final product. It might only be of interest in high performance, high cost low life
In a real engine the pressure ratio for design point speed at different altitudes will be different. Considering
the cycle analyses for condition 1, where the compressor pressure ratio is fixed at 2.25:1 for both AS and
SLS, the experimental data was used, and only ambient pressure corrections and air density changes were
made for SLS conditions. The effect of humidity was also not varied for the cycle even though it does have
environment.
In this chapter, the term, ‘module’, is used. A ‘module’ in Concepts NREC Axial design code is defined as an
IGV and the axial flow turbine rotor. Not to be confused, Concepts NREC Axial one dimension design code,
is a sub-component of the Concepts NREC design suite. Similarly, Concepts NREC AxCent three
dimensional design module is also a subcomponent of the Concepts NREC design suite and not the
‘module’ as referred to explicitly in the section 5.3 and section 5.4. Terminology for an axial flow turbine
The NGV and turbine rotor supplied for 3D scanning were used to determine the NGV blade aspect ratios as
well as the rotor blade aspect ratio and the hub-tip ratio. The new design is intended to fit in the specified
machine and there was no sense in recalculating these from the manufacturer’s existing design. Turbine
diameter is 66 mm, mean radius for the turbine and NGV is 27,95 mm.
For turbine design the specifications should include the pressure and the temperature values at the inlet and
exit of the stage. The design mass flow rate and the temperature drop across the turbine is sufficient to
determine the power the turbine will produce at a given speed and this is then used to baseline the power
specified for the turbine based on a complete compressor map. Velocity diagrams at the mean radius must
be chosen using the guidelines outlined in [1]. The number of blades in the rotor and stator must be
determined along with blade shape and spacing. This is done by calculating the optimum solidity for the
required flow deflection (turning) at the design point speed. [1], [32], [37]
The principle of the optimum solidity theory is that the number of blades must be a minimum value to turn the
flow and minimizing the frictional losses of too many blades whilst avoiding the flow separation from having
too few blades with too much incidence. Optimum solidity is related to blade aerodynamic loading. We use
the Zweifel coefficient (CL = 0.8) which minimises loss if the tangential lift coefficient CL is set at a constant
value of 0.8. Higher values of CL have been implemented, CL values of 0.9 and 1 on large turbojets are
common but these values are not frequently used in micro turbojets [32].
Figure 5-2: Profile losses in relation to solidity ratio and blade setting [32]
The tangential lift coefficient for a turbine cascade, such as that in Figure 5-3 is given by equation 5-1.
Figure 5-3: Diagram for calculation of tangential lift on turbine cascade of blades
[32]
𝑠 𝑊1 sin 𝛼1
𝐶𝐿 = 2 ( ) (cos 𝛼2 2 ) [( ) − tan 𝛼2 ]
𝑏 𝑊2 cos 𝛼1
where s is the blade pitch, b is the blade axial chord, W1 is the inlet relative velocity and W2 is the relative
velocity at the blade trailing edge. But the design approach used here is for a flared annulus and therefore
constant axial velocity. The implication is that the axial velocity Cx is then given by: [32]
𝐶𝑥 = 𝑊1 cos 𝛼1 = 𝑊2 cos 𝛼2
𝑠
𝐶𝐿 = 2 ( ) (cos 𝛼2 2 )[tan 𝛼1 − tan 𝛼2 ]
𝑏
If the optimum is then set to 0.8, the optimum axial solidity is given by equation:
𝑏
( ) = |2.5(cos 𝛼2 )2 [tan 𝛼1 − tan 𝛼2 ]|
𝑠 𝑜𝑝𝑡
For optimum solidity with the Zweifel lift coefficient (which implies CL = 0.8, we have the preliminary
𝛼1 is set at zero for no swirl at the inlet to the IGV and 𝛼2 = −56.7° using Equation 5-2
Number of vanes is 15
𝑈 = 347.46𝑚𝑠 −1
𝑠 = 5 𝑚𝑚 and 𝑏 = 7,685 𝑚𝑚
Number of blades is 23
According to [32], leading edge radii (Rle) typically fall between 0.05 s and 0.1 s (s, pitch).
Rle =0.6 mm
Trailing edge radii for the rotor and stator typically fall between 0.015 c and 0.05 c but for the preliminary
design for both the IGV and rotor the trailing edge thickness is set at (0.25 micron), essentially knife edged.
As flow stream lines approach the leading edge of the blade, they turn because of flow circulation that forms
when generating lift on the profile. The implication of this is that if a blade is not adequately cambered the
leading edge may produce an upstream throat that does not lie on the trailing edge of the blade. To
determine this induced incidence angle, the following equation is used [32]:
𝛼1 𝑐
∆𝜃𝑖𝑛𝑑 = 14 (1 − ) + 9 (1.8 − )
70° 𝑠
For the rotor design the ‘optimum’ solidity is 1.537 with the flow inlet angle at 33.3 °, hence using correlations
found by Dunavant and Erwin [33], the approximate induced incidence is 9.5 °.
The blade row stagger angle (λ) needs to be determined. This variable aids in defining the overall blade
shape and more importantly, it determines the shape of the flow passage. For a subsonic turbine, whose
throat is at the trailing edge of the blade, the guidelines laid out in [34] were used and the results of which
Figure 5-5: Stagger angle correlations for various inlet and exit angles by [34]
For an inlet flow angle of 33.3° and exit angle of 59.03°, a stagger angle (𝜆) = 27° is determined using
Figure 5-5. This would be evaluated during the detailed design of a blade passage because it is an iterative
process which requires a constant evaluation of boundary layer formation, passage flow stability and overall
losses through the passage. With the aid of CFD some of these effects may be evaluated but this is very
much a heuristic approach for which there is no defined optimization method or discrete optimum value [32].
Making use of the ‘setting angle’ approximation, an estimation of the blade chord c may be made by solving
𝑏
cos 𝜆 =
𝑐
Rather than having pure ‘straight back’ blades, it is beneficial in subsonic turbines to have some degree of
curvature, e, (see Figure 5-1) on the suction surface of the blade upstream of the throat to ensure the flow
downstream at the throat remains attached. This is beneficial because accelerating flows have favourable
boundary layers with lower resultant losses both aerodynamically and in terms of heat transfer coefficient. A
𝑠
range of shape values that give favourable blades is 0.25 < ( ) < 0.625 [32].
𝑒
𝑠
Selecting a mean value of ( ) = 0.4375 in the favourable range, 𝑒 = 11.428 𝑚𝑚.
𝑒
Calculation of the throat size requires that the throat Mach number be chosen and this value is specified to
be 0.7. The flow outlet angles may now be estimated using 2 equations that define flow between Mach 0 and
0.5 and then flow at Mach 1. The throat Mach number calculation of flow exit angle for Mach 0.7 will be
7 𝑜 𝑠
|𝛼2 |0<𝑀𝑡<0.5 = [ (|cos−1 ( )| − 10°) + 4° ( )]
6 𝑠 𝑒
and,
𝑜 𝑠 (1.786+4.128(𝑠⁄𝑒)) 𝑜
|𝛼2 |𝑀𝑡=1 = |cos −1 ( ) |− ( ) | sin−1 ( )|
𝑠 𝑒 𝑠
According to [32] these may be written as a single expression which predicts the flow angle at Mach
And substituting them into 5-9 we have a predicted exit flow angle at Mach 0.7 of,
𝑜
cos|𝛼2 |𝑀𝑡=0.7 =
𝑠
We have 𝑜 ≅ 3𝑚𝑚 at the mean radius of the rotor blade [1], [32].
Designing for an expansion ratio of 1.7, the density at station 2 of Figure 5-6 is given by:
𝑝2
𝜌𝑠 =
𝑅𝑇2
WithT1≈T2 =1150 K in micro turbines, ρ2 =0.866 kg/m3 and the annulus area may be calculated at station 2
by:
𝑚̇
𝐴2 =
𝜌2 𝐶𝑎2
From the constant stage due to the flared annulus, Ca2=277.9 ms-1 and A2 = 0.0013 m2. The throat area is
𝐴2𝑁 = 𝐴2 cos|𝛼2 |
A2N =0.000707 m2. These calculations are repeated for station 1 and station 3 of Figure 5-6. According to the
theory in [1], the annulus radius ratio range to consider is 1,2-1,4. A summary of calculated information for
Station 1 2 3
The mean radius (rm) for the blade design is 27.95 mm, the hub radius (rr) is 21.46 mm and the tip radius (rt)
is 34.33mm. Tip clearance which is designed for when cold is 0.5 mm this reduces to 0.22 mm at 1150 K
with Inconel 600 blade material, this is calculated using the method in section 5.6.
The Concepts NREC software environment completes these calculations at multiple locations along the span
of a blade design. The software is then used to build a 3D blade model, lofted from the hub to the shroud.
The summary of all the hand calculations are presented in Table 5-2
Table 5-2: Summary of hand calculations at mean radius and design condition
The terminology and modes explicitly described in this section are as named within the Concepts NREC
Axial design code input section. The point of doing this is so that if desired, the design can be reproduced
with ease.
The first inputs required by the user include defining the machine type in the package Concepts NREC
Axial® as well as other general module parameters. This information describes the type of axial flow
machine the user wishes to begin designing and limits the scope of the computations to the relevant
requirements. The NREC Axial® software package deals exclusively with axial flow turbomachines. These
may be pure reaction, pure impulse or a combination of the 2 types. The component type in the ‘module’
must then be defined as an IGV, rotor, stator, turbine or compressor. For the design that follows here,
The analysis mode then needs to be defined. For the results and outputs obtained here, the mode
‘Geo(P0T0AngPr)’ was selected. This mode allows the user to input static temperature (T0) and static
pressure (P0) with flow angle at the inlet, total-to-static pressure ratio, mass flow rate, outer reference radius
as well as flow and loading coefficients. The general module input is for the upstream station, this selects
local far upstream inlet conditions. For this design the mode selected was ‘LOCAL_UPSTREAM’, this implies
that the input conditions are at the inlet of the first component, the IGV.
The working fluid passing through the module (‘module’ as defined in sections 5.3 and section 5.4) needs to
be defined. The type selected is Concepts NREC own model for ‘real combustion gases’. The selection is
‘combust.spr’ within the code. Finally the machine rotational speed is set by the user for the design point.
The flow entering the module needs to be defined in terms of the flow type at the inlet in the absolute
reference frame. It’s useful at the first attempt to assume subsonic flow entering the NGV because
combustors require low to intermediate subsonic flow speeds to maintain the combustion flame front. The
subsonic flow assumption will be used to resolve the velocity triangles at the inlet of the NGV. The flow
velocity profile approaching the NGV is defined by in-built correlations. The available selections include
uniform and non-uniform flow conditions at the inlet to the module. For the first iteration, the selection
parameter is the ‘HT_CONSTANT’ profile. This section implies that the flow is uniform at the inlet to the NGV
and only the mid-span input data is required. Specifying the mid-span data allows the program to calculate
the hub and shroud conditions. Fluid mass flow rate is set in kg/s as well as inlet static temperature and inlet
static pressure. The default blockage value is zero, but a suitable range for axial flow turbomachines is 0-
0.17. For the first iteration, a blockage of 0 % is used, later this was revised to around 20 %.
In this section the user sets the objective values for the design target of the first module. First the predesign
mode needs to be selected. The mode for this iteration is ’FLOW+LOAD_COEF’, with this the number of
stages will be computed and the loading coefficient corrected to meet the integer value for the NGV by input
The next parameter to define is the geometry mode which controls geometry generation during the
preliminary design phase. The selection made is ‘GEO_DEFAULT’. This minimises user control over blade
geometry by making use of built-in correlations from experimental data collected at Concepts NREC. The
design flow path is set using the parameter ‘R (m)_CONST’ which sets the design flow path at the constant
mean radius which was prescribed earlier by the ‘HT_CONSTANT’ selection profile.
Blade configuration is defined next, ‘VAR_SECTIONS’ selection is made to allow for the development of the
blades such that the blade profile is variable from hub to shroud. The stage configuration is set as
‘FULL_STAGE_M’ which computes stage with the rotor and stator included with an increased rotor hub
loading. The allowed mean radius is user defined in millimetres (mm) by means of the parameter
‘Rad(m)ref_des’ and is set at 27.95 mm which is the maximum radius that is permissible for the turbine
The scale coefficient for the axial blade chord is set at a value of 5.8. This parameter directly controls the
axial chords which are estimated with the built-in correlations in Concepts NREC design code and it is used
to control the number of blades and blade loading. Stall limit is also affected by this control when working
with axial flow compressors. The default pre-set for the coefficient of area ‘Coef_Area’ is introduced to
compensate for 3D flow effects in the preliminary design with the default value being 1 and the acceptable
range from 0.95-1.05. Like the coefficient of area, the coefficient of reaction ‘Coef_Reaction’ also
compensates for 3D effects of flow but this parameter controls the slope of the hub and shroud line contour
by redistributing the reaction between the rotor and the stator component. The default value is 1 and the
For the first iteration design the flow coefficient is set at 0.6 by the parameter ‘Coef_Flow_des, this is defined
by the meridional velocity upstream of the rotor. Typical values for axial flow turbine stator and rotor stages
range from 0.3-1.2 depending on blade loading. The average load coefficient designated by the input
‘Coef_Load_des’ for the design is set at a value 1.38 based on the optimum value formula designed to
deliver a 50 % reaction for the machine at the mean line of the design. The general equation for any reaction
𝟎.𝟓
𝑪𝒐𝒆𝒇.𝑳𝒐𝒂𝒅_𝒅𝒆𝒔 = (𝟒(𝒇𝒍𝒐𝒘 𝒄𝒐𝒆𝒇𝒇𝒊𝒄𝒊𝒆𝒏𝒕𝟐 ) + 𝟐 + 𝟒(𝑹(𝑹 − 𝟏)))
The average flow angles at the inlet and at the exit of the machine are defined by the parameters
‘Angle_des.in’ and for the exit ‘Angle_des.out’ in meridional degrees. The default value set by the program is
zero degrees. This implies axial flow far upstream at the inlet and axial flow with no swirl far down stream of
the rotor exit. The total-to-static pressure ratio is set at 1.7. This is the borderline turbine expansion ratio and
directly controls the power delivery of the turbine. The design is based on an expected isentropic efficiency
Ambient reference conditions specified are 101.3 kPa and standard temperature of 291.15 K. The partial
admission reference is set at 1. This parameter controls the local downstream admission mode when
‘PA_MODE~E’ is set to scaled operation (‘PA_SCALED’) for the module components. Default convergence
settings are used. These are 100 iterations with a tolerance of 0.001. Greater convergence could be used
but the solution accuracy does not increase sufficiently to warrant the additional computation time. Rotational
speed for the module is set at 100 000 RPM maximum, design point speed is 95 000 RPM.
The first iteration 2D design has a prismatic blade, see Figure 5-7. This implies that the blade aerodynamic
loading is not uniform from hub to shroud because the blade tip speed is higher than the hub speed. To
evenly aerodynamically load a blade, the blade should be staggered from hub to shroud to accommodate the
speed differential at the hub and shroud. This is commonly referred to as blade twist.
The effects of uneven blade loading results in the flow potentially choking at the shroud line trailing edge of
the rotor blade passage. Loading the blades unevenly makes mechanical design of the rotor hub problematic
as stress fluctuations may initiate unpredictable flapping modes in the blades as well as the disk as a result
of this. It is for this reason that prismatic blades are not common in turbomachinery.
In some low cost, hobby grade micro turbojets, prismatic blades are used. This is because manufacturing
cost and complexity are minimised in hobby grade turbojets. It is significantly more complex to cast and
machine twisted blades in comparison with prismatic blades. In high performance applications, like UAV and
Figure 5-8: Prismatic blade meridional profile with mean line blade section in Axial
In Figure 5-8 the vertical axis is the components (NGV and rotor) radius from the axis or rotation. The
horizontal ‘Z-axis’ gives an indication of component width in millimetres. The blue component is the stator
and the red is the turbine rotor blade. Below each component, there is a profile view of the blade taken at the
mean line, since the blade is prismatic, this profile doesn’t change from hub to shroud. The blade
Mass-flow Rate
0.325 0.331 0.352
(kg/s)
Total-to-Total
0.811 0.766 0.635
Efficiency
Turbine inlet
1150 1150 1150
Temperature (K)
Turbine inlet
300 300 300
Pressure (kPa)
At design point the turbine produces 37.46 kW of power. This is with a modest expansion ratio of 1.68:1.
Typically, small axial flow turbines have expansion ratios around 2:1 per stage in multi-stage arrangements
[1]. To sufficiently stay away from stage loading that would result in 2:1 expansion, the ratio of 1.7:1 was the
design objective. The reason for this is the gas turning would be excessive (more than 70 degrees in the
turbine) in an attempt to achieve the required pressure drop for the power output of a single stage.
Incorporating blade twist into the design uniformly loads the turbine blade, see Figure 5-9. Due to the better
aerodynamic profile from hub to shroud secondary flow effects will also be reduced. The 3D rendering is
Figure 5-9: Twisted blade merdidional profile and blade section profiles hub to
shroud in Axial
The green profile section in the meridional view is the mean line, the blue profile is the hub profile and the
red is the tip profile for the design. These sections in the stator and rotor show the extent to which the blades
are twisted. It also is clear to see that for a twisted rotor design to work effectively, it needs to be coupled
with a stator that can effectively turn the flow. One should not simply incorporate blade twist in the rotor, the
effect of twist would be nullified as the inlet flow angles at the leading edge of the rotor would not be useful
The 3D rendering of the rotor is presented in Figure 5-10. The hub in the rendering is not contoured, this will
be completed later. A complete 3D view of the stage will be presented in the next section based on AxCent
analysis.
Total-to-Total
0.852 0.839 0.761
Efficiency
Turbine inlet
1150 1150 1150
Temperature (K)
The power produced in the twisted design at design point mass flow rate is 42,22 kW, see Table 5-4. The
expansion ratio is marginally higher than that of the prismatic blade design at 169:1. This is sufficiently close
to the design target of 1.7:1 for the stage. The aerofoil family used in both the prismatic and twisted design is
This program is used for detailed 3D geometric design and for rapid flow analysis of multi-staged axial and
radial turbomachinery through a streamline curvature technique. AxCent employs powerful features for axial,
mixed-flow, and radial geometries that can be combined to design almost any turbomachinery. AxCent
includes several options for real-time interactive flow analysis utilizing a number of CFD solvers. AxCent
uses an axisymmetric through-flow solver for rapid analysis of axial flow in compressors and turbines using
industry-standard loss and deviation models but this does not necessarily apply to micro turbojets, this type
Figure 5-11: Detailed meridional view of stator and rotor with flared annulus
geometry
The flow passage in Figure 5-11 is flared which is important in that the design was such that the axial flow
velocity is consistent through the rotor. This is only achieved with a flared annulus. The flared annulus
accommodates the decrease in density of the gas as it expands through the rotor.
Figure 5-12 is a complete stage rendering of the rotor and IGV. Maximum thickness at the hub is at 40%
chord and at the shroud maximum thickness at 50% span. The thickness distribution is more biased to a
higher loading from 50% chord to the hub. Although blade twist was used to reduce tip loading on the rotor,
the reduced load has been shifted toward 30% - 50% chord with a narrower passage throat closer to the
hub. The flow will thus be at a higher velocity in this area and CFD will be required to assess the sonic
Parameter Value
∆𝑻 (K) 139
The blade points for the IGV and the turbine rotor are in Appendix B. The points are in Cartesian
coordinates. For the rotor, the blade span was divided into 9 discrete planes. The blade profile was divided
into pressure and suction surfaces, each made up of 100 points. Concepts NREC code optimized the design
for throat Mach number of 0.75. This was a default value that could not be changed by the user.
Table 5-6: Concepts NREC optimized blade geometry for a throat Mach number
0.75
The hand calculations from section 5.2 for the mean radius with a throat Mach number of 0.7 yielded the
Table 5-7: Calculated mean line blade angles for throat Mach number 0.7
There is very close correlation between the mean line hand calculations of blade angles from Table 5-2 with
those calculated by the NREC code in Table 5-7. NREC Design CFD Evaluation
5.4.2 Background
Having produced software based design, it is worth assessing the performance prediction with computational
fluid dynamics. Where performance testing is not possible for a prototype, CFD lends itself as a tool for
performance evaluation at low cost whilst knowing it is not a substitute for experimental work on a prototype.
Concepts NREC AxCent has a ‘push-button’ CFD package which is allows for a seamless transfer of 3D
geometry and flow settings into a comprehensive turbomachinery orientated CFD package. The Pushbutton
CFD generates results based on full 3D Navier-Stokes equations. Mesh generation capabilities of the
Pushbutton CFD as will be applied in this thesis are classic H-grid and advanced topology OH-grid.
5.4.3 Aim
The reason for conducting CFD through the stage design is to determine the expansion ratio, to assess the
sonic conditions in the flow through the stator and the rotor and finally to obtain a base line performance
level which may be used during a future prototype experimental performance evaluation.
Some useful information that may be acquired includes a look into the extent of flow separation from hub to
shroud. The flow incidence at the leading edge of the turbine rotor will be determined as this may initiate the
Axcent input conditions are generated in Axial from mean line calculations. AxCent is used to recalculate
Inlet swirl at 0 % span (hub) is zero degrees. At 100 % span (shroud) inlet swirl is calculated to be 7.88
degrees. Inlet temperature is 1150 K, inlet pressure is set at 300 kPa. Momentum equations are solved by
specifying the inlet mass flow, an alternative to this approach is to specify the inlet total pressure. Inlet
energy is prescribed using inlet temperature rather inlet enthalpy. Absolute angles are used when specifying
At the exit, the pressure is specified as opposed to mass flow. This specification is determined by a rapid
blade-to-blade solver for Quasi-orthogonal (QO) components at maximum volume flow rate up to theoretical
predicted choking in the flow field through the geometry. The QO volume flow rates indicate that no areas in
the blade-to-blade field are choked. This rapid solution is run at the initiation of the CFD run.
The conditions in Table 5-8 represent the design data objective values. In assessing the CFD results, these
data will be revised with the results generated. Of interest from the CFD solution is the static temperature of
the flow leaving the rotor, Mach number of the flow leaving the rotor as well as the fluid pressure leaving the
rotor.
inlet to the stator to the rotor exit with a partial extension for CFD. An OH-grid is a hybrid mesh that has
characteristics of a wraparound O-grid, with the same characteristics of a shapeless H-grid. Cell generation
in the area where the boundary layer will develop in Figure 5-14 appears sufficient for the solution to be
moderately representative and inclusive of downstream turbulence development. The multi-block solver was
used. For the single passage a mesh of 104 118 cells was produced, this is sufficient for a single passage.
Ideally a mesh as dense as possible is desired but computing resource use increases exponentially. This
was the densest mesh with the maximum number of clearance cells possible on the machine available at the
University.
Figure 5-14: Shroud line mesh topology illustrating boundary layer cell resolution
0,093%. Y+ was 57, whilst a finer mesh would have been ideal, hardware limitations at the University
prevented this, optimal Y+ values in the tip would be closer to Y+ of 12. Overall, the flow entering the stage
had inlet static pressure mixed out and mass averaged at 279,88 kPa in Figure 5-15.
The static pressure plot of the flow leaving the rotor in Figure 5-16 suggests a mixed out static pressure of
196,72 kPa.
The static pressure field through the mid-span in Figure 5-17 indicates a static pressure of 185.47 kPa. This
specific regions of low pressure in the field suggests that a separation close to the hub surface has occurred
of flow separation. The Mach number at mid-span in the throat of the stator is approximately Mach 0.75.
a total-to-static temperature drop across the stage of 105 K. The high temperature at the hub and shroud
locations for the NGV is likely due to the boundary layer that has developed and as such the heat transfer
out of the NGV hub surface and shroud surface is lower as there is little turbulent mixing.
Judging by the CFD results collected in Table 5-9, it is clear that the turbine CFD result is not too far away
from the objective from the mean line design. However, proper experimental work is required to validate the
CFD. In terms of mass flow rate, the CFD suggests a 2.91 % increase in mass flow is likely with the specified
conditions. However, the turbine expansion ratio is lower than is expected. This can be attributed to the
degree of separation in the turbine but also to the lower than expected temperature drop across the stage.
Bulk efficiency is lower than design by 2.2 %. It is clear that a refinement of the design in order to minimise
the separation might be necessary. Alternatively, a refinement of the CFD mesh and a change in grid type
may also produce a better result. It is important to note that a mesh dependence study was not conducted
When predicting the performance of a turbine stage, losses need to be predicted and properly understood.
Typically losses are determined during cascade testing. Losses in turbines may be broken down into 3
During the design in this thesis, loss compensation is limited to minimizing tip clearance because this has a
documented effect on secondary flows and possible stalling downstream of the turbine rotor.
Group 1 losses:
These are losses in pressure of the working fluid, usually due to boundary layer thickness and friction.
Group 2 losses:
Increases in enthalpy due to heat addition from friction in the turbine effectively lowers the temperature drop
across a turbine and subsequently this reduces the power extracted from the flow. Group 3 losses:
Group 3 losses:
These can be attributed to windage loss due to friction effects in the viscous working fluid. This group of
losses is related more directly to the type of working fluid and is not directly in the designers control, however
The turbine casing in a micro turbojet is a thin walled cylinder, an area which may well be approximated to
that of a thin circular ring when taken in cross section [8]. The circumference of the thin circular ring is given
by:
𝐶0 = 2𝜋𝑟0
where C0 is the initial circumference of the ring at 21°C and r0 is the radius at 21°C. The incremental change
𝑑𝐶 = 𝐶1 − 𝐶0
𝑑𝐶 = 2𝜋𝑟0 𝛼𝑑𝑡
where α is the coefficient of linear thermal expansion for the material with units (mm/°C) and dt is the change
in temperature. Representing Equation 5-17 in terms of diameter, the following equation is obtained [8]:
𝑑1 = 𝑑0 (𝑑𝑡𝛼 + 1)
With this equation the expansion of the rotor shaft, turbine blades and turbine case can be approximated. d1
is the diameter of the rotor after thermal expansion, d0 is the initial diameter of the rotor. To make an
assessment of the running clearance of the Inconel 600 turbine at design point temperature of 580°C a
substitution of the coefficient of thermal expansion for Inconel 600 is made into the 5-18. The results show
that at 21°C the rotor clearance is 0.74mm and at 580°C it is reduced to 0.53mm.
This calculation does not take into account dynamic oscillations of the rotor. Measurement of dynamic tip
clearance is beyond the capability of the test rig at the University of Johannesburg. The small size of the
rotor and the complexity of measurement of dynamic fluctuations in tip clearance make this an important
variable to determine, dynamic clearance could be measured or estimated, the size of the gap could be
5.7.1 Background
The turbine hub supports and transfers the aerodynamic loads to the central shaft and has to resist
deformation due to centrifugal force. Finite element analysis (FEA) was conducted on the rotor hub design to
assess the level of the transferred loads in SolidWorks® 2011. For FEA to be meaningful, dynamic operating
conditions must be realistic. Material mechanical properties for Inconel 600 must be considered during the
FEA. Only the rotor disk was considered here. Inclusion of the blade geometry, as would be expected in a
turbine casting of this nature, was not possible due to Solidworks 2011 meshing utility being unable to
generate a mesh on the blade surfaces. This was investigated further and it became clear that only a
dedicated FEA software tool, was able to generate a multi-grid mesh, which allows a conventional mesh to
be generated on the hub and a fillet-compatible mesh on the blades. A dedicated FEA code was not
with interpolated interfaces. The material density for Inconel 600 is 8 190 kg/m3. Aerodynamic loads are
calculated in Concepts NREC Axial design code and are not possible in Solidworks.
When running the stress FEA for Inconel 600, the exact material specification was not available in
SolidWorks 2011. An Inconel 625 alloy was used. The density of Inconel 625 alloy is within 3 % that of the
Inconel 600 with identical Poisson’s ratio and tensile stress characteristics. The FEA will be conducted at
12 500 rad/s (120 000 RPM). The FEA resulted in von Mises stress of 225 MPa near the shaft bore for the
The Concepts NREC design code estimates a stress of 371 MPa this is inclusive of aerodynamic loading
which has an axial and radial component of force as well as centrifugal stress. The hub design in the
Concepts NREC code is for an arbitrary un-contoured disk of the appropriate hub radius. Refinement to the
design first evaluated in Solidworks has resulted in a reduction in the disk stress only. This was done by
increasing the radius of the rotor face contour in the region of the shaft bore. For Inconel 625 the disk stress
was 219 MPa at 12 500 rad/s, no aerodynamic loads were included. This was achieved by changing fillet
The node resolution, see Figure 5-21, of the mesh is sufficient and evenly distributed from the hub centre-
line to the outer radius. Refining the mesh further did not produce any significant change in results, cell count
was 114 277 cells on the surface. This was the densest mesh possible for the cell type on the machine
available at the University. The centrifugal stress on the disk spinning in a vacuum would be 219MPa, in an
operating environment for the turbine, with aerodynamic and thermal loads, the peak stress predicted by
Figure 5-22: FEA von Mises stress distribution in the impeller design
The stress distribution for the Inconel 625 alloy is depicted in Figure 5-22. This may be used for future work
when prototype rotors are developed. Working drawings for the rotor geometry are included in the Appendix.
The position of the peak stress is close to the bore. In some small high speed turbine rotors, a boreless
design is incorporated to maintain maximum material in the region of peak stress to reduce stress
concentration.
Using the maximum and minimum stress from the von Mises stress distribution from the FEA on the rotor
disk, the approximate life of the rotor may be approximated using the S-N data for various different materials.
This method is useful when predicting the life of a component before a prototype has been made and tested.
Calculation of the simplified Goodman equation is required to obtain an R= -1 equivalent stress which is then
used to give an indication of the number of load cycles till failure [36].
𝐹𝑈𝑇 ∙ 𝜎𝑎
𝜎𝑅=−1 =
𝐹𝑈𝑇 − 𝜎𝑚
The amplitude of the cyclic loading has a major effect on the fatigue performance The S-N relationship is
determined for single specific loading amplitude obtained by the Goodman equation. The amplitude is
expressed as the R-ratio value, which is the minimum peak stress divided by the maximum peak stress [36].
𝜎𝑚𝑖𝑛
𝑅=
𝜎𝑚𝑎𝑥
For 𝜎𝑚𝑖𝑛 = 23.4 𝑀𝑃𝑎 and 𝜎𝑚𝑎𝑥 = 219.1 𝑀𝑃𝑎, we have that 𝑅 = 0.107
The stress values are obtained from the FEA results shown in Figure 5-22.
𝜎𝑚𝑎𝑥 + 𝜎𝑚𝑖𝑛
𝜎𝑚 =
2
𝜎𝑚𝑎𝑥 − 𝜎𝑚𝑖𝑛
𝜎𝑎 =
2
Since turbine life is only usefully estimated when at operating temperature, the ultimate tensile stress FUT
must be used at temperatures that the turbine would operate at. Inconel 600 has a tensile strength of
approximately 800 MPa between 500 ℃ and 760 ℃. As a result, the solution to the simplified Goodman
This result is applied to a modified S-N curve for Inconel 625 (figure 4a) from the work in [35].This has had
the experimental data for standard Inconel 625 plate specimens extrapolated as an approximation to an
exponentially decreasing gradient toward the minimum graph value of 200 MPa. Since 115.33 MPa is lower
than the y-axis minimum, it can be said with relative confidence that the rotor design would survive in excess
Operating at a maximum speed of 100 000 RPM, the rotor would complete more than 108 cycles in
approximately 16.67 hours. Considering that the design point speed is 95 000 RPM and only in off-design
mode would the rotor reach 100 000 RPM, this time may be sufficient for low endurance operations. Typical
hobby grade micro turbojets are completely overhauled after 50 hours of design point operation. In many
instances, turbine rotor and compressor impellers are replacement items after 50 hours.
5.9 Discussion
Concepts NREC design code uses empirical correlations to account for tip losses in turbomachines. These
correlations are augmented with experimental work conducted at the Concepts NREC headquarters in White
River, Vermont, USA. These values are characteristic of turbines with higher aspect ratio blades, typically
higher than 3. The loss models are unlikely to accurately account for the extent of losses on very short micro
turbojet blades. The loss models determined for large mass flow engines had to be used as the design
models for small turbojet engine. Relying on the impact that a small leakage flow has in a full scale engine
does not produce realistic results in terms of leakage losses and the effects of leakage on secondary flows.
Hand calculation results for the mean line blade geometry appear consistent with the results obtained in
Concepts NREC Axial mean line design code. The benefits of the code itself are that it is able to perform
multiple calculations from hub to shroud, allowing optimum blade pitch and throat conditions along the full
blade span. This is highly beneficial for the design of micro turbojets where radial loss gradients are higher
than in larger engines. Since the micro turbojet is finding greater application in modern aviation, it will prove
beneficial to seek added efficiency by exploiting the predictive strengths of both CFD and meanline theory.
5.10 Conclusions
Detailed design analysis of the gas turbine was conducted following experimental bench tests of an existing
gas turbojet engine. The losses associated with small scale turbomachinery were effectively represented in
Gasturb 12 and in doing so the results of Chapter 4 may hold more value as they are not based on empirical
correlations for large gas turbines exclusively. When using the experimentally measured losses as an input
to the Gasturb cycle code after testing in Chapter 3, there was a close correlation between the experimental
work and the code result in terms of pressures, temperatures and thrust.
From the test data it was observed that the combustor pressure drop is greater than is typical for a turbojet
engine and was approximately 5.4 %. This is to be expected because combustor pressure losses being
greater in micro turbojet combustors than in larger annular combustors which is typically only 3 %. Also
noted, the compressor pressure ratio is relatively low and the turbine inlet temperature is low, principally a
consequence of the lower than design engine RPM and the inability to employ blade cooling in such small
blades. The engine did not reach the design point speed of 95 000 RPM but did reach 91 740 RPM. It is
worth noting that there is no accurate method to control turbine inlet temperature on the test rig. The engine
control system was set to ensure conservative operation of the fuel pump. Due to safety concerns, it was not
advisable to make changes to the fuel pump ramping curves and as a consequence slight drift in TIT was
encountered. Generally, TIT is increased by increasing fuel flow, this allows the engine RPM to climb. At the
An important conclusion to draw is that the experimentally measured data and the simulated results in
Gasturb have an acceptable correlation. The effects of combustor pressure drop and tip leakage in the
turbine rotor are difficult to quantify for small micro gas turbines. What has been shown is that the effects of
these losses will be predictable in a cycle code optimized for large turbojet engines provided component
Thermodynamic improvements to an engine turbine design will only be verified if prototype turbine rotors are
manufactured for testing after rigorous CFD analyses with a dedicated turbomachinery CFD code such as
NUMECA Fine/Turbo. This was not available at the University due to cost. The tests would be ideally
A concluding summary of data as is relevant to the turbine stage design, the inlet temperature and pressure
for the Concepts NREC design, CFD analysis of the design, experimental test results on the engine test
bench and Gasturb 12 approximation of testing results is presented in the Table 5-10.
Table 5-10: Summary of turbine design data, CFD data, rig test data for the turbine
(EGT)769
The NREC design was done with TIT of 1150 K. The reason for this was to operate a mixed flow
compressor at pressure ratio of 3,1 with mass flow rate of 0,325 kg/s, requires a minimum TIT of
1150 K if thrust is to be 120 N or greater. The mixed flow compressor was to be designed as an
upgrade to the test engine by another post graduate student at the University.
The test rig did not reach design point speed of 95 000 RPM. It was suspected that the fuel pump
was not capable of supplying a higher fuel flow rate due to engine manufacturer fuel flow limits set in
the controller to keep TIT below 600 °C. Maximum engine speed achieved was 91 740 RPM.
Mean line aerodynamic modelling of the rotor blade by hand calculations and by Concepts NREC
software produced very similar results to one another, see Table 5-2 and Table 5-6.
Cycle analysis using Gasturb software applying experimental data as an input produced similar
results to the experimental work thus verifying the cycle analysis conducted in Gasturb.
Experimentally measured pressure drop across the turbine was 7,28 % different to the NREC
design. However, the NREC design is done at sea level static conditions. The test rig is at 1750 m
above sea level in Johannesburg, South Africa where ambient pressure is 12,54 % lower than sea
level. Test results and the design results are consistent and within realistic expectations of the
practical turbojet cycle when correction for ambient conditions is made to the design cycle
performance.
6 Recommendations
6.1 General
Ideally the turbine design in Concepts NREC Axial and AxCent would move onto the detail design,
manufacture and engine integration phase but only after a complete and fully defined CFD analysis followed
by a full rotordynamic analysis for stress and vibration. A prototype would be tested and the results
compared again to the simulations in cycle code, CFD and the experimental data. This will further validate
the extent of the losses that are expected in practical applications as well as the cross correlation of results
and data from experiments and simulation in software. Ideally to match the turbine design to a suitable
compressor, a compressor map would be required, so that the final turbine design would be matched to the
compressor performance.
Change fuel ramping values in the controller to exceed the manufacturers EGT limitations to attain
Strip down the entire engine and get detailed scans of the as-built components to model the
6.3 Design
In terms of the design, the following is recommended:
Design work should place additional emphasis on off-design operation as well as design point
Investigation into methods for increasing turbine inlet temperatures that extend beyond turbine rotor
design into combustor design and material selection such as thermal barrier coatings.
A fully characterized compressor map should be provided to the turbine designer to ensure that the
turbine design delivers sufficient power and does not force the compressor into surge or stall,
especially during off-design operation. This would be incorporated with a detailed combustor model
A thorough CFD analysis in a dedicated CFD code for turbomachinery, such as NUMECA
A turbine rotor test rig should be designed to test turbine rotors with real combustion gas properties.
Investigation into thermal barrier coatings and their uses in radial flow turbines and axial flow
Design of an integrally cast rotor and shaft could be proposed, stress analysis on a “bore-less”
References
[1] H.I.H. Saravanamuttoo, G.F.C. Rogers, H. Cohen, Gas Turbine Theory 6th Edition, Pearson Prentice-
Hall 2009
[2] L. Moroz, P. Pagur, Y. Govorushchenko, K. Grebennik, SoftIn Way, Paper delivered at Int. Symp. On
Heat Transfer in Gas Turbine Systems, Comparison of Counter-Rotating and Traditional Axial Aircraft
Low-Pressure Turbines Integral Detailled Performances, 9-14 August 2009, Antalya, Turkey
[3] Rolls Royce Technical Publications Department, The Jet Engine 5th Edition, Crapaci Aeronautical
Library, Renault Printing Co. Ltd, 1986 (revised 1996) ISBN 0902121235
[4] P. M. Boyce Ph.D PE, Gas Turbine Engineering Handbook, 2nd Edition, Gulf Professional Publishing,
[5] A. Giampaolo, Gas Turbine Handbook – Principles and Practice, 3rd Edition, TFP Inc. CRC Press, 2006,
ISBN 0881735167
[6] K. G. Budinaki, M. K. Budinski, Engineering Materials: Properties and Selection, 8th Edition,
Pearson/Prentice-Hall, 2005
[7] D. W. Richerson, Historical View of Addressing the Challenges of Use of Ceramic Components in Gas
Turbine Engines, ASME Turbo Expo: Power for Land and Sea, Barcelona, Spain 2006
[8] J. E. Shigley, C. R. Mischke, Mechanical Engineering Design, 6th Metric Edition, International Edition,
McGraw-Hill, 2003
[9] Sonntag, Borgnakke, van Wylen, Fundamentals of Thermodynamics, 6th Edition, Wiley, 2003
[10] H. Barker, AGARD Conference Proceedings no. 276, “Ceramics for Turbine Applications” (1979) 24/1-
[11] G. Ziegler, J. Heinrich, G. Wötting, Review relationships between processing, microstructure and
properties of dense and reaction-bonded silicon nitride, Journal of Material Science #22, pg. 3041-4086,
1986
[12] W. D. Carruthers, P. F. Becher, M. K. Ferber, J. Pollinger, M. van Roode, Proceedings of ASME Turbo
Expo 2002, Advances in the Development of Silicon Nitride and Other Ceramics, GT-2002-30504, June
[13] O.E. Balje, Turbomachines, Design, Selection and Theory, Wiley International, 1981
[14] P. P. Walsh, P. Fletcher, Gas Turbine Performance 2nd Edition, Blackwell Publishing 2004 ISBN 0-632-
06434-X
[15] Dixon Chandley, Use of Gamma Titanium Aluminide for automotive Engine Valves, President Metal
Casting Technology
[16] Maki et al, Development of a high performance TiAl Engine valve, SAE 1996.
[17] Titanium Aluminide Machining, Schmite-Cincinnati Millicron – February 12, 1995 Symposium.
[18] A.S. Rangwala, Turbo-Machinery Dynamics Design and Operation, McGraw-Hill Mechanical
Engineering, 2005
[19] J.D. Mattingly, W.H. Heiser, D.T. Pratt, Aircraft Engine Design 2nd Edition, AIAA Education Series,
[20] B. Sunden, M. Faghri, Heta Transfer in Gas Turbines, Developments in Heat Transfer, Witpress
[21] N. Cumpsty, Jet Propulsion, Cambridge engine Technology Series:2, Cambridge University Press, 1997
Hill, 2004
[23] A. W. Judge, The Testing of High Speed Internal Combustion Engines 4th Edition, London Chapman &
[24] V.V. Raghavan, High Temperature Corrosion Failure of Super Alloy Turbine Blades, Dukhan, Qatar,
Technical White-paper
[25] T.A. Wallace, R.K. Bird, Development of Oxidation Protection Coatings for Gamma Titanium Aluminide
Alloys, S.N. Sankaran, Lockheed Martin Space Operations, VA, NASA Langley Research Centre VA
[26] R.G. Wing, I.R. McGill, The Protection of Gas Turbine Blades, A Platinum Aluminide Diffusion Coating,
Rolls Royce Limited and Group Research Centre, Johnson Matthey and Co. Limited
[27] M. Smarsly, N. Zheng et. Al. MTU Aero Engines, Advanced High temperature Turbine Seals Materials
and Designs.
[28] D. You, M. Wang et. Al. Effects of Tip-Gap size on Tip-Leakage Flow in Turbomachinery Cascades,
[29] C. Bell, The design of turbine internal cooling systems, Process management, Engineering Design
[30] J. Ling, K.C. Wong, S. Armfield, Numerical Investigation of a Small Gas turbine Compressor, School of
Aerospace, Mechanical and Mechatronic Engineering, University of Sydney, New South Wales
Australia, 2006
[31] Layton, Edwin T. "From Rule of Thumb to Scientific Engineering: James B. Francis and the Invention of
the Francis Turbine," NLA Monograph Series. Stony Brook, NY: Research Foundation of the State
[32] D.G. Wilson, The Design of High-Efficiency Turbomachinery and Gas Turbines, The MIT Press
[33] J.C. Dunavant, J.R. Erwin, Investigation of Related Series of Turbine-Blade Profiles in Cascade,
[34] S.C. Kacker, U. Okapuu, A mean line prediction method for axial flow turbine efficiency, J. Eng. Power
properties using standard and micro-specimens of base materials Inconel 625, Inconel 718 and Ti-6Al-
[36] D. Japikse, C. Osborne, W.D. Marscher, M.J.Platt, Structural Analysis of Centrifugal Impellers,
[37] S. L. Dixon, Thermodynamics of Turbomachinery, 3rd Edition, University of Liverpool, Pergamon Press,
1978
[38] R. H. Aungier, Turbine Aerodynamics – Axial-Flow and Radial-Inflow Turbine Design and Analysis,
ASME Press, New York, 2006
[39] J. H. Horlock, Axial Flow Turbines – Fluid Mechanics and Thermodynamics, Robert E. Krieger
Publishing Company, Huntington New York, 1973
In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if material had to be removed,
a note will indicate the deletion.
ProQuest 28281832
Published by ProQuest LLC ( 2021 ). Copyright of the Dissertation is held by the Author.
ProQuest LLC
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, MI 48106 - 1346